Sei sulla pagina 1di 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 978-1-4614-9151-4ISBN 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
Publishers 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 bachelors and masters 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 60,000lb. and move
with life-like dexterity, which motivated me to return to school and earn a masters 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 universities. 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 ordinary 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 packages 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 energetic properties of translational and rotational mechanical elements systems, illustrated by the
step response of first-order systems, are presented in Chap.4. Chapter5 emphasizes the analogies between fluid, electric, and thermal elements. Chapter6 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.
Chapter7 reviews the fundamentals of linear algebra, and introduces state-space. Mathcads and MATLABs RungeKutta 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 RungeKutta 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 those who have. Chapter9 introduces block diagrams and block diagram algebra, the basic concepts of closed-loop feedback control, the s-plane, and partial fraction expansion. Chapter10
presents frequency response. Chapter11 considers AC circuit analysis as a specific application
of frequency response. Chapter11 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 individual 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 mousearm.
Lafayette College

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

Contents

1 Introduction to System Dynamics 1


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

viii

Contents

2.4Superposition  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 Eulers Equations  64
2.7.4 Complex Exponential Form Z e j  65
2.7.5 Rotating Complex Exponential Unit Vector  68
2.7.6Example: Solution of a Homogeneous Equation using Complex
Exponential Unit Vectors  71
2.8Solved Problems Illustrating the Method of Undetermined Coefficients  74
2.8.1Example Problem One: First-Order Step Response  74
2.8.2Example Problem Two: Non-Oscillatory Second-Order
Step Response  77
2.8.3Example Problem Three: Oscillatory Second-Order
Step Response  79
2.9Eigenvalues and Response Characterization  84
2.9.1 First-Order Step Responses  84
2.9.2Non-Oscillatory Second-Order Step Responses  86
2.9.3 Oscillatory Second-Order Step Responses  88
2.9.4 Time Step for Response Calculations  89
2.10Laplace 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
3Introduction to the Linear Graph Method, Step Responses,
and Superposition 117
3.1Introduction  117
3.2 Introduction to the Linear Graph Method  117
3.2.1 Energetic Model  118
3.2.2Newtonian Formulation of the Force-Mass-Damper Model  121
3.2.3Linear Graph Formulation of the Force-Mass-Damper Model  121
3.2.4Examples Illustrating the Linear Graph Method  129
3.2.5Summary of the Introduction to Linear Graphs  135
3.3 The Heaviside Unit Step Function  136
3.3.1Differentiation 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
3.5.1Categorization of First-Order Step Responses  144
3.5.2Step Response of a First-Order Mechanical System  145
3.6 Time Shift  147
3.7Superposition of Heaviside Step Functions  148
3.8Superposition of First-Order Step Responses  149
3.8.1Scaling, Time Shifting, and Superposing Unit Step Responses  151

Contents

ix

3.9Superposition of Second-Order Step Responses  154


3.9.1Overdamped or Non-Oscillatory Step Response  154
3.9.2Underdamped or Oscillatory Step Response  156
3.10 Initial Condition of Energized Systems  161
3.10.1Example of Second-Order Pulse Response
of an Energized System  162
3.10.2 Initial Value Method  165
3.11 Solved Problems  169
3.11.1Step Responses of Initially De-energized System  169
3.11.2Step Responses of Initially Energized System  171
3.11.3Second-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
4.1Translational 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
4.2.4Damper: Viscous Friction Energy Dissipation  210
4.2.5 Translational Mechanical Sources  213
4.2.6Summary  216
4.3 The Sign Problem of Mechanical Systems  216
4.4Drawing Linear Graphs from Mechanical Schematics  219
4.4.1 Linear Graph Symbols  219
4.4.2Force Source Acting on a Parallel Mass-Damper System  219
4.4.3Force Source Acting on a System of a Mass and Two Dampers  221
4.4.4Force Source Acting on System of a Mass and Two Dampers  221
4.4.5Force 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
4.6.2Mass Moment of Inertia of Primitive Shapes  226
4.6.3Mass Moment of Inertia Calculated from Area Moment of Inertia  226
4.6.4Mass 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
4.7.1Components with a Single Unknown Energetic Parameter  237
4.7.2Components 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

Contents

4.8.4 Springs in Series  257


4.8.5Equivalent 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
5.2.3Pressure-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
5.4Calculating Fluid Element Parameters from Fluid Properties
and Geometry  282
5.4.1 Fluid Resistance  282
5.4.2 Fluid Inertance  283
5.4.3Fluid Capacitance (Hydraulic Accumulators)  284
5.5Fluid 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
5.6.1Analogies 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
5.7.2Electrical Capacitance: Energy Stored in an Electric Field  290
5.7.3Electrical 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
5.8.2Modes of Heat Transfer and Their Corresponding
Thermal Resistances  295
5.8.3 Thermal System Elements  297
5.8.4 Thermal Systems  299
5.9Equivalent 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
5.9.4Fluid 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 and Ferromagnetic Materials  319

Contents

xi

 agnetic Flux Density B, Flux , and Applied Magnetic


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

xii

7Vector-Matrix Algebra and the State-Space Representation


of Dynamic Systems  411
7.1Overview  411
7.2 Vector-Matrix Algebra  411
7.2.1 Matrix Addition  412
7.2.2 Matrix Multiplication  413
7.3Operating on a Vector-Matrix Expression with a Linear Operator  414
7.3.1Laplace 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
7.6State-Space Representation of Dynamic Systems  422
7.6.1 State Variables  422
7.6.2Example 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
7.6.6Vector-Matrix Form of the Output Equations  427
7.6.7Numerical Solution of the State and Output Equations  427
7.7Example Derivations of State and Output Equations  428
7.7.1Third-Order Dynamic System Example: A Rotational Mechanical
System  428
7.7.2Fourth-Order Dynamic System Example: A Spring-Mass-Damper
System  430
7.7.3Fourth-Order Dynamic System Example: A Fluid-Mechanical
System  433
7.8 Why State-Space is called State-Space  437
7.9Expression of Systems Equations in State-Space  438
7.9.1Algorithm to Express a Higher-Order System Equation
Without Differentiation of the Input as State Equations  438
7.9.2Algorithm to Express a Higher-Order System Equation
with Differentiation of the Input as State Equations  440
7.10 Eigenvalues and Eigenvectors  443
7.10.1Eigenvalues  443
7.10.2Eigenvectors  444
Summary  444
Problems  445
Chapter 7 Appendix  457
Mathcads RungeKutta 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
MATLABs RungeKutta 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

Contents

xiii

8 Finite Difference Methods and MATLAB  467


8.1Finite Difference Approximation of Differential Equations  467
8.2Euler Method, Forward Stepping, Finite Difference Algorithm  468
8.2.1MATLAB Programming of the Euler Method, First-Order System  469
8.2.2Euler Method Solution of Second-Order State Equations  471
8.2.3Example: 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 RungeKutta Method  479
8.4.1Two-State, Fourth-Order RungeKutta Algorithm and Code  479
8.4.2Three-State, Fourth-Order RungeKutta Algorithm and Code  480
8.5Programming Non-Linearities and Input Functions  481
8.5.1 Common Non-Linearities  482
8.5.2Input Functions, Non-Linearities and the RungeKutta 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
MATLABs 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
9.1.4Differential Equations and Transfer Functions as Linear Operators 521
9.2Laplace-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
9.3.3Differentiation, 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
9.4Time-Domain Block Diagrams of Differential System Equations  535
9.4.1Block Diagram Without Differentiation of the Input  535
9.4.2Block Diagram with Differentiation of the Input  537
9.5State Equations as a Time-Domain Block Diagram  542
9.5.1Drawing a Block Diagram of Existing State Equation  542
9.5.2Drawing a Block Diagram from the Energetic Equations  542
9.5.3Drawing a Block Diagram of State Equations from
a System Equation  543
9.6 The s-Plane  545
9.6.1Poles, Zeros, and Pole-Zero Transfer Function Form  546
9.6.2Stability  547
9.6.3 Real Component , the Decay Rate  549
9.6.4Imaginary Component , the Observed, Damped Frequency  549
9.6.5 Damping Ratio  549
9.6.6s-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
10.1Overview of Sinusoidal Excitation and Frequency Response  577
10.1.1Calculating Magnitude and Phase Angle from a Response  580
10.1.2Transient and Steady-State Response of a First-Order System  582
10.1.3Transient and Steady-State Response of a Second-Order System  584
10.2 Frequency Response Relationship  588
10.2.1Example Frequency Response Calculation  588
10.3Derivation of the Frequency Response Equation  590
10.4Fourier Series Approximation of Periodic Signals  593
10.5 Bode Plots  595
10.5.1 Review of the Properties of Logarithms  595
10.5.2Decibels  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
10.6Asymptotic Approximation of the Log-Magnitude Bode Plot  605
10.6.1 Asymptotic Approximation Form  607
10.6.2Asymptotic Approximation of First-Order Factors  608
10.6.3Asymptotic Approximation of Underdamped
Second-Order Factors  610
10.6.4 Integrator and Differentiator Factors  613

Contents

xv

10.6.5 Gain Factors  614


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

xvi

Contents

11.11.1NEMA  656
11.11.2US 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

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 boundary, between storage modes within the system, and is dissipated as heat by friction, electrical 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 substitution, 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.

1.1Introduction
Websters New World Dictionary defines dynamic as relating to energy or a physical force in motion: opposed to
staticrelating to or tending toward change. System
is defined as a set or arrangement of things so related or
connected as to form a unity or organic whole: as a solar system, irrigation system, supply system. Our focus will be on
physical systems within mechanical engineering, specifically, the machines and processes which mechanical engineers
design and maintain. The changes we are interested in are
changes to a systems internal physical power variables,
such as force, velocity, pressure, fluid flow rate, current,
voltage, temperature, and heat flow rate. Our objective will
be to predict the dynamic response of systems comprising
mechanical, fluid power, and electrical elements.
Energy is a quantity which flows. Dynamical systems
change due to the flow of energy across the boundary separating the system from the surrounding environment, and the
flow of energy between elements within the system. Energy
flows into a system from a power source, such as a battery,
a pump, a moving belt, or a spinning shaft. The amount of
energy which flows from the source is determined by the
response of the system to the supply of energy.
Oscillations, such as pressure fluctuations and mechanical vibrations, are created by energy flowing between energy
storage elements within a system. When mechanical systems

vibrate, energy flows back and forth between the kinetic energy of motion and the potential energy of elastic deformation. What we will model as energy storage elements are,
in fact, properties of physical objects. Often, a single physical object will have multiple energetic properties. We will
represent each property in our system models as a different
element, Fig.1.1.
Eventually, oscillations die down, as the flowing energy is
dissipated as heat by friction, viscous shear, electrical resistance, or magnetic hysteresis. System dynamics does not include the study of heat engines. Heat engines merit the entire
discipline of thermodynamics. Consequently, energy dissipated as heat is not converted back into other forms of energy.
It must remain as heat. However, for a system to function
indefinitely, the energy dissipated as heat must be removed
from the system, or the systems temperature will rise to damaging levels. Therefore, although system dynamics does not
include thermo, it includes the basics of heat transfer.
Dynamical systems change over time. We will work with
time scales which range from microseconds to hours. Regardless of the time scale, we will define time t=0 to be the
instant an input or action is applied to a system. We will then
divide time into three periods relative to time t=0: before,
shortly after, and long after. The latter two are more formally
known as the transient time period, and the steady-state time
period. The changes of a systems power variables during the
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 Introduction to System Dynamics

2
Fig. 1.1 An energetic system
consists of interacting energy
storage, dissipation, and conversion elements separated from the
surrounding environment by the
system boundary

System boundary

in
Energy flow into the
system from the source

Energy
storage
element

Energy
storage
element

Energy
dissipation
element

out
the input is applied, (2) the type, or functional form, of the
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 energy source. The systems change due to the functional form
or shape of the input, such as step change F (t ) = F0 u s (t ) or a
sinusoid, F (t ) = F0 sin(t ), is termed as the forced response.
The forced response corresponds to the particular solution
of a differential equation, as you may recall from your study
of differential equations. In the ideal models we will create, the forced response will exist from the instant after the
input is applied, at a time t = 0, and forever after (i.e., t = ).
The response of a system to change in its energy source is
termed as the natural or unforced response. The natural
response occurs whenever the system is disturbed from
its current energetic equilibrium by the application of a new
input. A systems natural or unforced response is always the
same shape though it differs in amplitude. It corresponds to
the homogeneous solution of a differential equation. The natural response dies out with time, as the energy flows due to
the disturbance dissipates, and the system approaches a new
energetic equilibrium. When the transient period ends, the
amount of energy flowing into the system equals the amount
flowing out. The system is then said to be in steady-state.
Changes of the systems power variables during the transient period consist of the sum or superposition of the permanent forced response and the temporary natural response.
The shape or functional form of the natural response depends
only on the dynamic system and is independent of the shape
of the input function. We will be particularly interested in the
transient response of the system, which corresponds to the
startup or shutdown of real machines. It is during the transient period that the amount of energy stored in the system
changes and when the power variables, such as force, pressure, and electrical current, reach maximum or minimum
values. Consequently, the design values used to size the machine components often occur during the transient period.

Energy dissipated as heat


is lost from the system

1.1.1Why Study System Dynamics?


Why study system dynamics? The short answer is that mechanical design requires the ability to predict the transient
and steady-state response of energetic systems to arbitrary
inputs. A slightly longer answer is that most of the analyses
you have performed so far in your study of mechanical engineering have tacitly assumed that the system is in steadystate. In other words, you have ignored the transient response
of systems you have analyzed, and, accordingly, your analyses are incomplete.
The best answer to the question Why study system dynamics? is to prepare you for engineering practice. In many
ways, engineering practice is the inverse of the undergraduate study of engineering. As a practicing engineer, you will
understand the applicable engineering analyses. Your effort
will be focused on determining what is assumed to be known
already in textbook problems. Specifically, you will first
need to create an engineering model, which is a simplified
representation of the component or system to be analyzed.
Next, you must establish the material properties, the operating conditions, and the forces, torques, pressures, velocities,
flow rates, voltages, or currents acting on the component or
system. Only now you are ready to begin your analysis. The
study of system dynamics develops ones ability to create
engineering models. An understanding of system dynamics
is needed to establish unknown material properties from dynamic tests. Finally, the variables which act on individual
components within a system are the power variables of a
dynamical model of the system. System dynamics analyses
yield predictions of how the power variables in a system
vary with time, allowing the engineer to determine the critical loading for their analysis.
Many of the topics and analyses of system dynamics
appear in other subjects and applications. The allied topic
of mechanical vibrations preceded system dynamics in

1.1Introduction

historical development, and utilizes many powerful analytical techniques, based on Newtonian analyses, energy
methods, and virtual work. Electric circuits are dynamical
systems. The method we will develop in this course, the linear graph, is a network representation of energetic systems
which borrows heavily from electric circuit analysis. The
most direct application of system dynamics is to develop the
mathematical models of systems needed for automatic control theory. Automatic control is added to a system to change
its behavior. There are two fundamental types of automatic
control, Boolean or on-off control and closed-loop feedback
control. Boolean control sequences the operations of a machine. Boolean control is formulated in Boolean logic and
can make machines appear intelligent. Closed-loop feedback
control changes the dynamic response of a machine. The design of closed-loop feedback control requires a mathematical
model of the dynamic system to be controlled. The mathematical form of the model varies with the mathematical form
of the controller. Analog or continuous control design requires a differential system equation. Digital control requires
a difference equation. Modern or state-space based control
requires a set of first-order differential equations presented
in vector-matrix form. As daunting as these mathematical
terms may be on first reading, even the engineering student
most doubtful of their mathematical ability can master the
concepts and techniques needed to describe a dynamical system as a mathematical model.
A wide variety of engineering mathematics is needed to
analyze dynamic systems. Although everyone enrolled in
systems dynamics has passed differential equations, many
students have forgotten most of what they knew. Others may
believe that they never understood differential equations
even though they passed the course. To those who question
their capacity to understand engineering mathematics, take
heart in the often cited quote, You can only understand what
you already know. Accordingly, it is common for mechanical engineering students to pass the prerequisite course, differential equations, by learning the various solution methods
and then later gain an understanding of differential equations
with their subsequent use in system dynamics. A student
should not be embarrassed if this is their situation. He or
she should accept it and make the effort needed to gain an
understanding.
We will review the portion of differential equations which
we will use, presenting the solution of differential equations
from an engineering perspective. Further, we will restrict
ourselves to a limited number of techniques. Specifically,
we will emphasize that physical systems are capable of only
a limited number of responses to common inputs. Consequently, solving the differential system equation will be
performed by, first, identifying which type of response the
input to the physical system will produce and then sizing
the response, by scaling its amplitude and duration. Among

the benefits of this inputoutput or input-response perspective is that it allows practicing engineers to make rational
decisions based on their understanding of how a system can
and cannot behave when they lack enough information to
formulate the system equation, which is often the case.
The Laplace transformation, a topic in differential equations, will be reviewed and then used in a slightly different
way than students may have done in the past. We will introduce vector-matrix algebra to more easily represent the
complicated dynamic systems mathematically. This mathematics may be new to many students and will be presented
as such. Likewise, we need some mathematics from complex
analysis. This also will be new material to many students. A
limited understanding of complex analysis yields significant
benefits in computational tools, including complex impedances and phasors.
Numerical computation using Mathcad, MATLAB, or a
package of your preference is an integral part of systems dynamics. Even for the limited class of differential equations
we can solve analytically, evaluating the solution manually (i.e., with a calculator) to present the results as a plot
can be tedious and error prone. Automated computation is
much easier and readily provides graphical representation
of the inputresponse relationship. Further, we will exploit
the property of superposition of linear systems so that we
can superimpose, or sum, many individual, simple input and
response plots to predict the response of systems to complicated inputs. We will find that once the syntax and methods
of the computation software are familiar, developing MATLAB programs to approximate and solve differential equations using finite differences is quite straightforward and
will free us from the need to simplify important aspects of
a physical system to yield a system equation which can be
solved analytically.
Although students are most apprehensive about the mathematics of dynamics systems, it is a students ability with
engineering modeling and analyses which is the key to success in system dynamics. It is important for mechanical engineering students to recognize that their curriculum progresses from the compartmentalized, sophomoric courses,
which introduce basic concepts, to the upper-level courses in
which engineering concepts, principles, and knowledge from
previous and concurrent courses are assumed and used. This
necessary progression from narrow to broad course content
can be difficult for some students, particularly those who
adopt an algorithmic perspective in which they see, or more
accurately, prefer to see, each problem as having one, and
only one, unique solution. Such a limiting perspective is a
crippling error in judgment. We will emphasize the reality of
engineering practice by highlighting the similarities between
apparently different engineering analyses, and working some
analyses using two or more of complementary techniques
which we must choose between.

The further engineering students progress through their


course work, the more choice they have when presented
with open-ended problems in which there is an objective,
such as design an X to satisfy criteria Y, but no immediately identifiable analyses which the student is expected to
apply. It is natural, at first, for students to seek guidance of
the form, What do you want me to do? which instructors
typically, and annoyingly, answer by reiterating, I want you
to design an X to satisfy criteria Y. A more helpful response
may be I want you to identify the questions (design variables) involved in designing an X to satisfy criteria Y, and
then answer them (i.e., determine the design values).
The instructors response is evasive, because open-ended
design problems are fundamentally different from problems
which illustrate and reinforce specific algorithms. In practice, engineering analyses are performed to either make design decisions or to validate design decisions. The analyses
used at the beginning of a design should be simple and quick,
even at the expense of accuracy because the design is yet to
be defined. The final analyses will be the most precise since
they will be based on the final geometry of the design and the
final predicted design loads, material properties, and operating conditions.
Typically, engineers can choose among a number of methods which will yield the result they need to make a decision.
Often, the various analyses were developed from different
starting points, at different times, and for different applications, and only later found to overlap. This is the case in systems dynamics. The choice of analyses depends on the type
and nature of the system, the result needed, and the personal
preference of the engineer. When an engineer is dealing with
a novel and important design, then the overlapping but different analysis is a relief because it allows the engineer to
check the result. If there were two analyses available, the
engineer would use the easier of the first two because, depending on the result, it may need to be repeated with revised
design parameters or, perhaps, abandoned completely. Following the iterative design, the second, more difficult analysis would be performed as a check on the first.
System dynamics is a relatively late development within
mechanical engineering, dating to the 1940s. System dynamics was preceded by three very different theoretical approaches, which all yield a system equation or set of system
equations that relate the initial condition(s) and the input
variable(s) to the output variable(s) using different but complementary analyses and techniques. They are, in order of
historical development, the following:
Newtons Laws of Mechanics (1687);
Lagrange Energy Methods (1788) and Hamiltons Energy
Methods (1834); and, lastly,
Kirchhoffs Circuit Laws (1845).
Our focus will be on a network method, the linear graph
method, which is an extension of Kirchhoffs electric circuit

1 Introduction to System Dynamics

laws to mechanical, fluid mechanical, and hybrid systems,


such as electromechanical systems. The linear graph method
draws analogies between the modes of energy storage and
energy dissipation as heat in different energy domains (or
types), based on the manner in which the power variables
of an energy domain are measured. The name, linear graph,
refers to the dispensing of distinct symbols to represent different types of energetic elements, as used in electric schematics, and representing all elements as a line connecting
two nodes with an arrow head denoting the assumed positive
direction for the variable, which flows or acts through the
element. The parameter abbreviation, i.e., R for resistance,
and M for mass, is used to identify the type of element.
We will also apply Newtons method. You are already familiar with Newtons method, but are likely unfamiliar with
its application to the vibration (or oscillation) of a system
of masses, springs, and dampers. Dampers are mechanical
elements that dissipate kinetic energy into heat, by shearing
a viscous fluid or deforming a viscoelastic material, typically
an elastomer, which is a synthetic rubber. You will find
that Newtons method will have many mathematical similarities with the linear graph method, even though the approach
is fundamentally different.
Lagranges and Hamiltons energy methods provide a different perspective on dynamic system from the linear graph
method we will develop. Lagranges and Hamiltons energy
methods provide important insights and useful techniques
for the analysis of energetic systems. The student interested
in dynamic systems is advised to study them in the future.
However, all texts must have a limit, so this text does not
include them.

1.1.2Engineering Models
A model is a simplified representation of reality. An engineering model represents only one (or some) of the attributes
of the real object, system, or process. Engineering models
can be physical, analogous, or mathematical. An example of
a physical model is a model airplane, which may only be a
polymer shell. It is a scale model, or a replica of the geometry, of an actual airplane. This type of model would be used
in a wind tunnel to model the aerodynamic forces acting on
a full-sized airplane of the same shape. Alternatively, the
physical model of an airplane may be a scale model, which
has motors, moveable control surfaces, radio control, and the
capacity to fly. Clearly, these two different types of models
of the same airplane have different levels of realism and different uses. The same is true for engineering models. The
level of realism varies with the intended use, which depends
on the stage of the design process the model is used in. In all
cases, an engineering model is a simplified representation of
the real entity.

1.1Introduction
Fig. 1.2 Creation of an engineering model of an actual physical
object or system is a filtering
process, in which only simplified
approximations and the relevant
properties are retained

Actual Physical
Object or System

Engineering models are abstractions of real, physical objects or systems. Abstract is a word, which intimidates students because one of its meanings is difficult to
understand, as in abstract mathematics. However, the
primary meaning of abstract comes from its Latin roots,
which means to pull or drag away from. Note that abstract shares a root with tractor and traction. Pulling
or extracting representative aspects from a whole to create a
simplified version is the meaning of abstract we will use.
This meaning will be familiar to some, since it is the same
usage of abstract as the abstract at the beginning of this
chapter, which is a short summary of it. Consequently, our
use of abstract will be close to simplified. For example,
if our objective is to predict the translational acceleration of
an object subjected to a force then only relevant property of
the object is its mass. If our objective is to predict the translational and rotational acceleration of the object, the relevant
properties are the geometry of the object and its density or,
if it varies, the distribution of its densities. It is important to
remember that every real object has mechanical, electrical,
and chemical properties. We often consider only one type or
class of properties at a time and forget that the real object has
all physical properties, some of which may affect the performance or reliability of our design.
All engineering analyses require an abstract model. Let
us deconstruct this last sentence. Although we commonly
employ the word, analyze, to mean work to understand,
the literal meaning of analyze is to take apart or break
into pieces, which is indeed the first step in understanding
anything complicated. Abstract can mean hard to understand, but it can also mean simplified, as in an abstract
painting. Model means a representation of something,
which includes only some of its attributes. Engineering
analyses are based on abstract models, since each physical
phenomenon which we must engineer for a part, machine,
or system typically requires a separate calculation. Although
the phenomena coexist in reality, we separate them, in order
to simplify the problem sufficiently, so that we can handle it
with the mathematical tools we have available. The model
used for a calculation depends on what we wish to know,
and contains only the properties and attributes of the corresponding physical object or system needed to describe that
phenomenon. When it is impossible to separate two different
phenomena because of their physical coupling, the engineering analysis is more difficult.

Modeling Process
Abstraction and
Approximation of
Relevant Properties

Engineering
Model

Although the term engineering model may be new, you


began using engineering models in your first engineering
course. What is new is that we are now considering the process of developing the model, Fig. 1.2. Previously, you were
presented a model as the starting point for learning a new
concept or computation. For example, in a statics problem,
the geometry of an object is retained, but the material properties of an object are simplified (or abstracted) by assuming
it is perfectly rigid. The assumption of perfect rigidity is
better termed the simplification of perfect rigidity since
it is a choice we make to simplify the engineering model.
How reasonable it may be to make the choice to simplify the
material properties by modeling the object as rigid is relative, and is determined by comparing the results computed
using the simplified model, with results computed using a
more realistic model, or, if available, with experimental data.
Needless to say, an engineering student first learning statics
cannot develop a more realistic model; nor does he or she
have a basis of experience to make the judgment of how reasonable an engineering model is. Consequently, engineering
modeling is not taught in introductory courses; the models
are simply presented.
Engineering computations are performed to provide
a basis for design decisions, and that design is a series of
choices. Early in the design process, few choices have been
made, and a great deal of uncertainty exists about the design,
the materials and manufacturing processes that will be used
to make it, and the operating environment it will be used
in. The initial engineering models must be correspondingly
simple. Often, a simple model is sufficient to reject or accept
a design choice. In other cases, the results fall into a gray
area and the initial, simple calculations are inconclusive, or
the cost or importance of the part merits refining the initial
results, by repeating the analyses on a more realistic and,
consequently, complicated model. For example, we may
first choose to model an object as perfectly rigid, in order to
consider only the static equilibrium of the object. Later, to
increase the realism of our model, we may choose to model
the material as elasticperfectly plastic.
The familiar elasticperfectly plastic material is a
model, Fig. 1.3. There is no material which actually behaves
like this, but it is a reasonable approximation for some important materials, and, consequently, is widely used. Although
the elasticperfectly plastic material model appears simple
since it consists of only two straight lines, it is a non-linear

1 Introduction to System Dynamics

Elastic
= E

1.1.4Extraction of Mathematical Statements


from Engineering Models

Plastic
= y

Fig. 1.3 Elasticperfectly plastic material model

material model and can greatly complicate engineering computations. We will return to this important point but, briefly,
a linear material (or property or constitutive) model is a
single straight line which passes through the origin.

1.1.3Mathematical Models
A mathematical model is an equation or set of equations
which form an inputoutput relationship, where the input
variables are known and the output variables are what we
want to know, Fig. 1.4. A mathematic model is based on an
engineering model. The various attributes and properties of
the engineering model can be described by simple mathematical statements. We work to create a compendium, or
list, of all equations which describe physical aspects of engineering model. These equations are then reduced to eliminate the unwanted variables, leaving just the input which we
know, the output we wish to calculate, and the independent
variables of time and position. For example, if we were sizing a part such as a shaft with a circular cross-section, the
properties would be the geometric constraints, such as the
required length and the maximum possible diameter, the approximate material properties of the shaft material, and the
approximate relationship between stress and strain (i.e., the
material model), the input would be the forces and torques
acting on the part, and the response would be stress, strain,
or deformation of the part.

Fig. 1.4 Mathematical engineering models are inputoutput


relationships used to facilitate
design decisions

The physical aspects of an engineering model, which can be


expressed as mathematical statements, clearly depend on the
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 devices, which are approximations and, therefore, never
completely true.
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
all circumstances. The accuracy of these equations is limited
only by our ability to measure the relevant quantities. Engineering models, on the other hand, are always approximate
and vary in accuracy from application to application. For
example, the linear elastic material model, = E , is quite
accurate for steel, before the stress reaches the yield point.
Conversely, = E is an approximation for aluminum,
which has no defined yield point.
To describe anything real with great accuracy is, at best,
difficult and costly in engineering time. It is often simply
impossible, because the information is not available, or the
variability between specimens or samples is so great, that
the uncertainty of the model makes its increased precision
meaningless. For example, most stress calculations are
performed assuming the material is ideally elastic, and the
stresses and strains are kept below yield to prevent permanent deformation. The material model for an elastic stress
calculation requires Youngs modulus E, a single datum. If
the stress will produce permanent plastic deformation, then
a non-linear relationship is needed. Typically, the ideally
elasticperfectly plastic model is used. It requires two values, Youngs modulus E, and either the yield stress, y, or
the yield strain, y. The yield stress or strain of a material is
not known with the certainty of Youngs modulus. Although
Youngs modulus for steel is independent of the specific
alloy and heat treatment, the yield stress and strain are not.
If the manufacturer or supplier does not provide test data for
these values, the designer must rely on nominal values found
in standards and handbooks. Nominal values of yield stress

Input Variable
Forces and Torques

Mathematical
Engineering
Model

Response or
Output Variable
Stresses, Strains,
or Deformation

1.1Introduction

and yield strain for a specific steel are not average values.
Although engineers often use nominal to mean normal,
nominal means in name only. In the case of yield stress,
a nominal value is often below the expected minimum yield
stress. This is a useful value if yield must be prevented, but
it is not the yield stress.
A more accurate description of a stressstrain relationship requires replacing the two straight lines of the elastic
perfectly plastic model with a single, continuous curve. This
type of non-linear model requires significantly more information than a piece-wise continuous model composed of
two straight lines. Non-linear stressstrain models are only
used in mechanical design where the economic value of the
improved design justifies the expense of the analysis, such
as the design of automobile bodies. Ultimately, the accuracy
of a material model depends on the variability of the material being modeled. Engineers deal with populations of
things, where the number in the population ranges from a
few to millions. There is always variation between individual specimens and samples drawn from the population. The
properties of populations are described stochastically using
probability to quantify the uncertainty. In order for a sophisticated non-linear material model to yield useful results, the
properties of the sample used to create the model must be
representative of the population of production parts. For example, plastic forming of steel changes it properties. How do
the properties of the steel in a location of an automotive body
correspond to the test specimens? In general, even the most
sophisticated engineering calculations are best described as
approximations.
The process of formulating a mathematical model from
an engineering model is aptly described as analysis, which,
again, strictly defined means to take apart. How one takes
an engineering model apart depends on whether the variables
of interest vary continuously with position and time, or only
with time. When materials are considered on the macroscopic
scale, i.e., we use a continuum model to represent materials with distributed properties. Systems in which the dynamic
response of variables change continuously with both position
and time, as is the case for pressure and velocity in fluids and
stress and strain in solids, are described by partial differential equations. Partial differential equations are significantly
more difficult to deal with than ordinary differential equations, which contain derivatives of only one variable (i.e., either time or position). In many circumstances, we can greatly
simplify a model, and avoid the necessity of partial differential equations with negligible loss in accuracy, by concentrating the distributed properties of a continuum into discrete
elements. The modeling process of representing distributed
material properties as discrete elements is known by the odd
term, lumping. A distributed property becomes a lumped
parameter. A familiar example of a lumped parameter is the
concentration of the mass of an object, at its center of mass.

The mass distributed throughout a volume is concentrated, or


lumped, as mass M at the mass center. Similarly, the elastic
strain of a spring is lumped into a spring rate or spring constant K.
The fact of the matter is that we must use very simple
equations to describe the physical aspects of the engineering model, or the resulting mathematical model is useless,
because we cannot solve it. By and large, most equations
which describe the physical aspects of an engineering model
are purely arithmeticeither sums, such as, an equilibrium
statement that the forces in the x-direction sum to zero, or
linear algebraic equations, such as = E . The most intimidating type of equation we will need to describe the physics of our models are linear, first-order ordinary differential
equations.

1.1.5Formulation of a Mathematical Model


from an Engineering Model
A mathematical model is derived from an engineering model,
by reducing the set of equations representing the physical
phenomena and properties of the engineering model to an
equation which yields the value of an output variable (what
we wish to know) as a function of the input variable (what
we do to the system) and the independent variable (usually
either time a position in space).
Reduction of the set of equations extracted from the engineering model to a mathematical model is conceptually the
easiest step of the process. We will reduce the set of equations to an inputoutput relationship using elimination by
substitution. We will begin, by choosing an equation from
the set of equations that contains the input variable, the output variable, or both. We will then substitute for a variable
in the equation, which is neither input nor output nor time.
Repeating this process eventually yields an equation whose
variables are only input, output, and time. The challenging aspect of the reduction is the number of substitutions
required to eliminate all of the unwanted variables. Simple
algebra can become error prone when there are a number of
steps unless done with care.

1.1.6Summary of Engineering Modeling


and Analysis
While each of your engineering courses presents new material, engineering analysis continues to build upon the same
framework, regardless of application. The commonality between your various engineering courses is quite deep and
permits understanding of the new through analogy to the old.
The method is summarized as follows:

1 Introduction to System Dynamics

8
Fig. 1.5a Model of a force
acting on a mass supported on a
frictionless surface. b Decomposition of the applied force into
horizontal and normal components

+x,+v

W = Mg

FN
Fx

Frictionless Surface

FR

1. Draw a picture(s) of what you have chosen to analyze.


This is your model. Annotate the picture(s) to define all of
the variables and parameters in your model.
2. Write mathematical statements of relevant physical truths.
3. Reduce the mathematical statements, eliminating all unknown variables, except the input and output variables,
and solve the equation (or system of equations) for the
output variable(s).
4. Evaluate the solution. Iterate if necessary.
We shall explore the analogies between the application of
engineering analysis to mechanical design and system dynamics in this chapter.

E W = Fx

(1.1)

where F is force and x is displacement. The dot or scalar


product is the product of the projection: the force vector in
the direction of the displacement vector, Fx, times the dis-

FR

placement vector x. It is only the component of force, F, in


the direction of the displacement, x, which does work. In
order for force F to do work on the mass the point of application of the force, F, must displace Fig.1.5. If the pointof
application of the force does not move, that is, if x = 0, then
the force does no work, and there is no energy transfer into
the mass.
Energy is a quantity which can be accumulated in a physical object, transferred from one object to another, and converted from one form of energy to another, until the energy
is ultimately dissipated as heat energy. Energy can flow. Further, given the opportunity, energy will flow. Power is the
flow rate of energy, expressed as

1.2Mechanical and Energetic Models


We will begin system dynamics, by reviewing the fundamental definitions of energy and power, as well as introducing concepts needed for energetic modeling of dynamic
systems. The flow of energy into, out of, and within physical
systems is what makes dynamic systems dynamic. Unfortunately, most of the physical variables mechanical engineers
work with cannot be directly observed. We must infer them,
by their effects which we can observe. For example, we cannot see force, but we can see the deformation of an object
subjected to a force, or the motion of a mass accelerated by a
force. Likewise, although we literally see light energy in the
form of photons striking our retina and feel heat, we cannot
sense any other forms of energy or power. We must infer the
amount of energy stored in an object by its deformation or
its motion.
A fundamental definition of energy is mechanical work

P=

dE
dt

(1.2)

Figure1.5a shows a force, F, acting on a mass supported by a


frictionless surface. This drawing is an energetic model of
a real, physical system. Engineering models are greatly simplified representations of real objects or systems. An energetic model is a simplified representation of just the significant
energetic properties or attributes of a system. The frictionless
surface in the model is an idealization, since there must be
some friction between two surfaces. By modeling the surface
as frictionless, we represent the friction force andthe energy
dissipating as heat, due to the friction between the mass and
the surface being negligible. Significant and negligible are
comparative adjectives that are always relative. In a Newtonian model, whether a force is significant or negligible is
judged relative to the other forces present. In an energetic
model, whether the amount of energy stored or dissipated is
significant or negligible is judged relative to the other quantities of energy in the system.
In Fig.1.5, the reaction force FR is approximately the same
magnitude as the applied force F. In a Newtonian model, the
reaction force FR would not be judged to be of negligible
magnitude. However, the reaction force FR is normal to the

1.2 Mechanical and Energetic Models

Round steel bar


Diameter D
Youngs modulus E
Yield strength y

F(t)

F(t)

F(t)

Non-Uniform
Stress

Round steel bar


Diameter D
Youngs modulus E
Yield strength y
Non-Uniform
Stress

= E

F(t)

Uniform Stress

L
Fig. 1.6 Axially loaded round steel bar

displacement of its point of application. The reaction force


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 idealization. What is not apparent is that identifying an element
of the energetic model as a mass is also an idealization.
Note that we described the element as being a mass rather
than having mass. This semantic subtlety restricted the element to the single physical property of mass or inertia, at
the exclusion of all other properties, such as stiffness, or its
inverse, compliance. Consequently, a mass is a mechanical
element that is perfectly rigid. If it were not, then the element would have also have compliance and either store or
dissipate energy due to deformation. Describing the mass as
a mechanical element further restricts it, by eliminating
all attributes of mass other than inertia, such as thermal capacitance. It is difficult at first to accept that the elements
of an energetic model describe real properties, but the elements themselves are idealizations. No real object has only
a single energetic property. However, many objects have a
dominant energetic property. For example, the springs of an
automobiles suspension have mass, electrical conductivity,
and thermal capacitance, but their dominant energetic property is strain energy storage. Conversely, it is not unusual
for a single real object to have multiple energetic properties
of approximately equal importance. When that is true, then
the single object is modeled by multiple energetic elements.
An ideal mass accelerated by a force is a dynamics problem.
Energy is stored in the mass as kinetic energy, the energy of
motion. When material properties are included in the model,
the motion of the point of application of a force is, in general,
the sum of displacement and deformation. The deformation
of a material can both store and dissipate energy, requiring
additional energetic elements in the model.
A model is a simplified approximation of the key aspects
or attributes of actual materials, objects, processes, and phenomena. Consider the familiar problem of calculating the
stress in a round rod due to an axial force F(t) using the
model shown in Fig.1.6. What assumptions and simplifications are stated or implied in Fig.1.6? In other words, what
physical models and constraints are used on the variables?
The callout provides two material parameters, Youngs
modulus E, and the yield strength y. These parameters de-

~D

~D

L
Fig. 1.7 Saint Venants principle for uniformity of stress. In an axially
loaded cylinder, non-uniform stress extends a minimum distance equal
to the diameter of the cylinder from the applied load

fine the model used to represent the material of the bar as


elasticperfectly plastic, Fig.1.3. The elasticperfectly plastic material model does not describe a real material. We shall
see that even this simplified model presents mathematical
difficulties because it is non-linear. It consists of two straight
lines, not one straight line. A linear relationship is a straight
line, which passes through the origin. A linear material
model allows scaling (doubling the force doubles the stress)
and superposition (sum of the stresses due to different forces
equals the stress due to the sum of the forces). The elastic
plastic model is only linear in the elastic region.
A second aspect of the model shown in Fig.1.6 is so
familiar, that it may escape notice. It is convention to represent the force acting on a physical object as a vector. Yes,
force is indeed a vector quantity. The vector representation
is physically accurate, but it is also abstract or simplified.
The modeling simplification is the omission of the details
as to how the force is applied to the object. The drawing
implies that the force is concentrated at a point which
certainly cannot be the case, since the stress at that point
would be infinite. Why then represent the force applied as a
vector? Models are simplifications, which contain only the
essence of the physical system needed to perform a calculation. A force applied to a machine element is represented
as a vector, when only the magnitude and direction of the
force are important. If an applied force is depicted as a vector, then how that force is generated and transmitted to the
bar is unimportant to that model or unknown.
To calculate the stress in the bar, we must assume that
stress is uniform on a transverse cross-section. The assumption of uniform stress sets a minimum distance from the concentration of stress of unknown magnitude where the force is
applied. Saint Venants principle is that the stresses are relatively uniform at a distance of approximately one diameter
from the applied force in an elastic object, Fig.1.7.
If the stresses are uniform anywhere over the length of
the bar, they will be uniform in the middle. A revised model
for the calculation of the stress in the bar which explicitly
indicates the location of the stress calculation is shown in
Fig.1.8.

1 Introduction to System Dynamics

10

F(t)

Non-Uniform
Stress

x1, v1

Round steel bar


Diameter D
Youngs modulus E
Yield strength y

F(t)

= E A(t)

x 2 , v2

F(t)

Translational Spring

Uniform Stress

~D

Fig. 1.10 Schematic symbol for a translational spring K showing the


locations of distinct values of translational displacement and velocity at
the ends of the element

L
_
2

Fig. 1.8 Revised model showing the cross-section on which the stress
is assumed tobe uniform

Elastic Deformation

F(t)

x1 , v1

Round steel bar


Diameter D
Youngs modulus E
Yield strength y

F(t)

x 2 , v2

Fig. 1.9 Axially loaded elastic bar of length L with elastic deformation L. Locations of distinct values of translational displacement and
velocity are denoted as x1,v1 and x2,v2

The last aspect of the model we will consider is the magnitude and mathematical functional form of the applied
force, F(t). We have noted that the vector representation of
the force applied to the bar omits the source of the force,
as well as the manner in which the force is transferred to
the bar, indicating that they are unimportant for the result of
the calculation. Are there physical limitations in the model
which constrain the magnitude of the force applied or how it
can vary with time? The magnitude is limited. The greatest
force which can be applied to the bar is that which causes
yield.
Fmax (t ) = y A
Hence, the operating range of the applied force is limited to
the following:
F (t ) y A

axial = E axial

(1.3)

axial

L
L

(1.4)

We can lump the distributed property of elasticity into a


single parameter by modeling the bar as a spring with the
elastic property, the spring constant, K. The deformation L
of the bar is the difference in the displacements x1 and x2 of
its two ends. The conventional notation for displacement is
the following: x1 x2 = x. We will use an alternative notation where the subject locations (or nodes) are identified by
subscripts. We will use the numbers, which identify the ends
of the element. Hence,


x1 x2 = x x12

(1.5)

We will use the same notation to denote the difference between velocities, pressures, voltages, and other analogous
variables, which vary across elements.
The elastic energy stored in the bar is recoverable. The
energy transferred to the bar to deform it plastically is dissipated as heat, as the metal shears and is not recoverable.
Although there are exceptions, such as head bolts in automobile engines and some other fasteners, the machine components are sized to keep stresses below the yield stress. When
stress is below yield, all of the mechanical work done on
the component is stored as elastic strain energy and can be
recovered. The schematic symbol for a significant amount
of strain energy storage in an energetic symbol is a spring,
Fig.1.10.
The familiar equation for a spring is


The bar shown in Fig.1.6 behaves as a spring, when its stress


is in the elastic region below yield, and the axial stress is
proportional to the axial strain


wherein,

FK = K x

(1.6)

where K is called the spring rate in machine design and the


spring constant in system dynamics. Although it will seem
awkward at first, we will use the notation of Eq.1.4 to express deformation of the spring element as the difference in
displacements of the two ends of the spring, denoted as locations or nodes one and two.


FK = K x12

(1.7)

1.3 Energy, Mechanical Power, and Coenergy

11

1.3Energy, Mechanical Power, and Coenergy

FK

1.3.1Strain Energy
Is there a dynamic constraint on the applied force, F(t), applied to the spring in Fig.1.10? Specifically, is it physically
possible to instantaneously apply an axial force, F(t), to an
elastic member? No, it is not. There are two complementary perspectives, both of which lead to the same conclusion.
The Newtonian perspective involves forces, displacements,
and their derivatives. If we were to posit a small but finite
force on a spring instantaneously, then we are saying that
the deformation of the spring due to that force can occur instantaneously. A finite deformation of the spring in zero time
requires infinite velocity of the ends of the spring, which is
physically impossible.
The complementary perspective is to view the spring as
an energy storage element, and look for similar physical
impossibilities. Recall that the differentials, d E and dt , are
infinitesimal. They are infinitely small, compared to the finite quantities, E and t. If there were a finite change of
energy E during an infinitesimal time, dt , then the power
P would be infinite.

dt

E
0

Instantaneous deformation would require infinite power and


is, therefore, physically impossible.
Work is defined as the dot product of a force and the
point of application of the force (Eq.1.1). Consequently,
there must be displacement of the point of application of a
force to transfer energy into or out of an element or system.
The energy transferred to a spring equals the work done on
thespring.

E Spring = WSpring =

FK dx12

(1.9)

We can dispense with the vector notation and the scalar product, when force and displacement are co-linear.

E Spring =

xmax

(1.10)

FK dx12

We can then eliminate either force or deformation from the


work integral, using the relationship between spring force
and deformation. Eliminating the spring force, FK, yields the
familiar expression for the energy stored in a spring

E Spring =

xmax

FK dx12 E Spring =

xmax

2
K= 12 Kx12

x12
Fig. 1.11 Elastic strain energy stored in a linearly elastic spring is the
area between the FK(x) line and the displacement axis

E Spring =

1
K x 2 max
2

(1.11)

A forcedisplacement plot of a linearly elastic spring is


shown in Fig.1.11 where the force, FK, is the force acting
through the spring, and x12 is the deformation of the spring,
which is the difference in displacement of the two ends of
the spring. The elastic strain energy is the area of this plot
below the curve.
Eliminating the deformation of the spring x12 yields a
less familiar, or, perhaps unfamiliar, expression for elastic
energy.

(1.8)

xmax

E Spring =

xmax

FK dx12 E Spring =

FK max

F
FK d K
K

Note that the upper limit of integration was changed from the
maximum displacement, xmax, to the maximum spring force,
Fmax. Evaluate the integral

E Spring =


1
K

FK max

FK dFK E Spring =

E Spring

1 2
=
FK max
2K

1 2
FK
2K

FK max
0

(1.12)

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


the spring is compressed or extended. The strain energy is
positive, because the spring force and the deformation have
the same sign in Eqs.1.6 and 1.7.
Newtonian analyses use the term, potential energy, for the
energy stored in a spring, because that energy has the potential of creating motion and being converted into kinetic
energy. The phrase, strain energy, will be used in this text to
differentiate it from other forms of stored energy, which can
be converted into kinetic energy.

1.3.2Mechanical Power
K x12 dx12

Real physical objects have multiple energetic properties.


Ideal energetic elements, such as an ideal mass, are abstract

1 Introduction to System Dynamics

12
x,v

Massless, rigid bar

F(t)

b1

quently, a transfer of energy. When we examine the second


term of Eq.1.14,

b2

Fig. 1.12 A bar modeled as massless and rigid acts as a force spreader in a mechanical schematic. The cross-sections of a piston in a cylinder are symbols for dampers which dissipate mechanical power as heat
by pumping fluid

simplifications with a single energetic property. An ideal


masss single energetic property is the ability to store kinetic energy. In order to have this single energetic property,
an ideal mass must be rigid. If an ideal mass were deformable, then it could store strain energy or dissipate energy as
heat through shear deformation, or both. It would have two
or three energetic properties, not one. An object modeled as
rigid, but not given the attribute of mass, can neither store
nor dissipate energy.
We include ideal rigid bars and rods in models of mechanical systems to transmit force from one energetic element to
another, Fig. 1.12. There must be displacement of the point
of application of a force to transfer energy into or out of an
element or system. If a rigid object is free to move, it can
transfer energy as well as force. Conversely, if a rigid object
cannot move, then it can provide a reaction force, but that
force can do no work. It follows that energy cannot be transferred into or out of a rigid object fixed in position. Motion
is required to transfer energy. There must be displacement
of the point of application of a force in the direction of the
force, in order to perform mechanical work.
From the definitions of work, Eq.1.1, and power, Eq.1.2:

P=

dE
dW d ( F x )

=
dt
dt
dt

(1.13)

The derivative of a dot product is evaluated as if the dot


product were conventional multiplication, yielding the sum
of two terms.


P = F

dx
dF
+ x
dt
dt

(1.14)

Examining the first term,


F

dx
dt

we find that the displacement is changing with time. There


is motion of the point of application of the force and, conse-

dF
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
to mechanical power.
= F
P

dx
dF
+ x
dt 
dt

Does No
Work

Mechanical power is the first term alone.




P = F

dx
dt

(1.15)

The time derivative of displacement x is velocity v:




dx
v
dt

(1.16)

Thus, mechanical power is the dot product of force and the


velocity of the point of application of the force.
(1.17)
P = Fv

1.3.3Strain Coenergy
We concluded in Sect.1.2.1 that it is physically impossible to
instantaneously apply an axial force F to an elastic member,
because a finite energy transfer in infinitesimal time requires
infinite power. We also noted that the deformation x of the
spring would occur instantaneously, which requires infinite
velocity. There is another perspective we can apply to that
question. If it were possible to instantaneously apply force F0
to an elastic member, such as a spring, and then create deformation x0, the work done on the spring would be the product.
W = F0 x0
This is the rectangular cross-hatched area in Fig.1.13.
We know that the strain energy that is actually stored
in the spring, when it is deformed x0 under force F0, is the
triangular area under the curve, Fig.1.11. The corresponding area above the curve is coenergy. Coenergy is a fictitious mathematical entity. It does not exist, but it is useful
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


amount, dFK. The expression,

F0
FK

Coenergy

Energy
x0

x12

Fig. 1.13 The work which would be done on a spring, if force F0 and
its resulting in deformation x0 could be applied instantaneously, is the
cross-hatch rectanglular area, F0x0. Only the work represented by the
triangular area below the FK(x) line is physically possible

Coenergy
Force, F

dF

x12 dFK dW

x(F)
F(x) Energy

also illustrated in Fig.1.14, does not represent an infinitesimal amount of mechanical work, because the displacement,
x12, does not change. There must be motion of the point of
application of a force for that force to perform mechanical
work. Therefore, we can conclude that integral of the infinitesimal product, Eq.1.19, is not work, but something else,
even though the quantity has the units of force times displacement. 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
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
spring rate K is a function of the spring displacement x, thus,

Displacement, x
Fig. 1.14 Energy is the area below the forcedisplacement curve. The
fictitious but useful quantity, coenergy, is the area above the curve

Non-linear springs can be designed to become either softer or stiffer with displacement, Fig.1.15b. For example, a
spring which becomes stiffer with displacement may have
the spring rate of Eq.1.21, plotted in Fig.1.14.

COE Spring

Fmax

x12 dFK

(1.18)

Although this integral resembles the intermediate result obtained deriving Eq.1.12, which expresses the amount of energy stored in a spring in terms of the force acting through
the spring, there is a subtle but important difference. Removing the derivative from spring displacement x and placing
it on the spring force FK indicates that the displacement is

Fig. 1.15a Non-linear spring rate,


K ( x) = 5,000 N/m + 50,000 N/m 2 x.
b Spring force vs. displacement
for the spring rate in Fig.1.15a

Energy and coenergy are equal, when there is a linear relationship between force and displacement. When the relationship between force and displacement is non-linear, then
knowing either the energy or the coenergy allows calculation
of the other, by subtracting it from the area of the rectangle,
FKmax xmax:
Energy = FKmax xmax

FK max

x( FK ) dFK

(1.22)

and
Coenergy = FKmax xmax

20

xmax

FK ( x) dx

(1.23)

5
4

15

FK(x) 3
kN

K(x)
kN
__ 10
m

5
0

(1.21)

K ( x) = K 0 + K1 x


to Eq.1.10, but with the derivative on the spring force, FK,
rather than on the spring displacement x.

(1.20)

F ( x12 ) = K ( x12 ) x12

dx

(1.19)

1
0

0.1

x12 , m

0.2

0.1

x12 , m

0.2

1 Introduction to System Dynamics

14

Elastic
= E

yield

Plastic
= y

yield

Stress,

Energy
density

Coenergy
Stress, density

Energy
density

Strain,
Fig. 1.16 Energy density is the area below a stressstrain curve. It is
the quantity of energy stored and/or dissipated deforming a unit volume
of material

Equations1.22 and 1.23 are Legendre transformations. A


Legendre transformation from thermodynamics is the relationship between enthalpy, internal energy, pressure, and
volume.
H = U + PV
(1.24)

1.3.4Energy Density
Strain energy and coenergy are expressed in terms of stress
and strain as energy density, which is the energy or coenergy per unit volume of material. The area below a stress
strain curve is the energy stored or dissipated in deforming a
unit volume of the material. Energy density must be integrated over the volume of a component to determine the amount
of mechanical work stored as strain energy and dissipated as
heat performing plastic deformation.
Calculating the energy density of the elasticperfectly
plastic material model shown in Fig.1.16 is an example of
the utility of coenergy. The energy density in the material is
calculated by subtracting the coenergy from the product of
the yield stress, y, and the maximum strain, max.


Energy Density = yield max Coenergy Density (1.25)

where the coenergy density is the triangular area between the


stressstrain curve and the stress axis, Fig.1.17.
Coenergy Density =


1
yield yield
2

(1.26)

Energy density is useful in understanding the rupture of materials and the behavior of materials during plastic forming,
such as bending, stamping, and forging. A phenomenon observed during bending metal is the recovery of some of the
bending deformation, when the bending force is removed

yield

Strain,

max

Fig. 1.17 Coenergy density is the area between the stressstrain curve
and the stress axis

from the work piece. This is termed spring back, and necessitates bending the work piece further than the desired
permanent bend. The mechanism of spring back can be seen
in Fig.1.17. The energy density of a metal plastically deformed is dissipated as heat and cannot contribute to spring
back. However, the material stores the elastic strain energy,
and that energy is recovered, when the bending force is removed, causing spring back.

1.3.5Kinetic Energy
An old joke among engineering professors is that students
part way through their engineering education believe they
understand energy but do not understand entropy. Their engineering education is complete, when they realize that they
dont understand energy either. Mass and energy are very
mysterious at both the atomic and subatomic scales, where
they are governed by quantum mechanics, and at extremely
high energies on the macroscopic scale, where relativity
governs. The transmutation of energy into mass and mass
into energy is a bizarre concept. However strange the implications of relativity may be, they are not as confounding
as the reality of quantum mechanics. It is easier to accept a
particle acquiring mass, as it is accelerated to great velocity,
than it is to accept that the same particle must also be viewed
as being a wave having no definite location in space, and
may spontaneously appear elsewhere. Although some phenomena mechanical engineers work with, such as electrical
conduction and radiative heat transfer, require quantum mechanics concepts to understand, most mechanical engineering is done on a human size and energy scale or not far from
it. Consequently, the effects of both quantum mechanics and
relativity are negligible, and we use Newtonian mechanics 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


actually understand. However, from the perspective of what
we can predict with confidence to make engineering decisions, we can describe the relationship between mass and
energy with Eq.1.1, the definition of energy as mechanical
work, and two additional algebraic equations, Eqs.1.27 and
1.28, and a differential equation, Eq.1.30.
Both mass and energy can move or flow. Although mass
and energy often move together, we account for them separately with two Conservation Laws, when we exclude atomic
reactions and relativistic velocities. Conservation of mass is
a statement that the difference between the mass that flows
into a system and that which flows out must be stored within
the system.


mass

In

mass

Out

= massStored

(1.27)

Conservation of energy is expressed similarly.




energyIn

energyOut = energyStored

(1.28)

In both Eq.1.27 and 1.28, we have defined flow into a system as positive and flow out of a system as negative.
The remaining equation needed to describe the relationship between mass and energy is Newtons Second Law, the
relationship between a force acting on a mass and its acceleration. We will express the familiar


F = Ma

(1.29)

as a differential equation
dv
F=M
(1.30)
dt
in order to write the equation in terms of velocity rather than
acceleration. Force and velocity are known as the power
variables of a translational mechanical system, since their
product is mechanical power, Eq.1.17.
Energy is a quantity that moves or, more familiarly, flows.
The flow rate of energy is power. Although it is common to
speak of a power flow, power is a flow rate. Power is the
energy flow rate into or out of an object, across a boundary,
or through a conductor or transmission element during a unit
of time. We will use both terms, power and power flow, since
the redundancy serves to remind us that power is a flow rate.
We will now investigate the relationship between Newtons Second Law and kinetic energy. If we restrict the motion of a mass M to a horizontal plane, so that it cannot gain
elevation against gravity, then the only type of energy the
mass can store is kinetic energy. We can eliminate the vector
notation, by assuming that the force acting on the mass and
the velocity of the mass are in the x-direction.

Fx = M

dvx
dt

(1.31)

Similarly, we can eliminate the dot product in the expression of mechanical power, Eq.1.17, when the force and the
displacement of its point of application are restricted so that
they are colinear. We again assume that force and displacement are in the x-direction:


P = F v P = Fx vx

(1.32)

We now substitute for Fx using Newtons Second Law:

P=M

dvx
vx
dt

Multiplying both sides by dt and rearranging the right side,

P dt = Mvx dvx
dE
using the definition of power P =
rearranged as
P dt = d E and then integrating yields dt

P dt = Mvx dvx d E = M vx dvx E = M

vx2
2

The result is more easily recognized as kinetic energy, when


rearranged as


E Kinetic =

1
Mvx2
2

(1.33)

Kinetic energy will be given the subscript M to identify it as


the energy of a mass. Although gravitational potential energy
is also stored in mass, we will find it convenient to express
that form of energy as work done against the gravitational
force, for reasons which we will discuss later.
1
(1.34)
E Kinetic E M = Mv 2
2
There is no such thing as negative mass, only accounting
deficits of mass or mass flow which yield negative signs.
We also generally think of energy as a positive quantity, but
here signs become a problem. Kinetic energy must be positive due to the square of velocity. However, the electrostatic
potential energy between two charges Q1 and Q2 separated
by a distance d can be either positive or negative, depending
on the signs of the charges. If both charges have the same
sign, then the force between them is repulsive, and the potential energy is defined to be positive. Conversely, if the
two charges are of opposite sign, then the force between
them is attractive, and the potential energy is defined to be
negative. The signs of potential energy required the same

1 Introduction to System Dynamics

16

arbitrariness of choice, as Benjamin Franklin made assigning


signs to charges. It will likely surprise you to learn that the
potential energy of a mass in a gravitation field is negative,
because gravity is an attractive force between masses. We
routinely ignore the negative sign, when we calculate what
we refer to as the potential energy of a mass raised from the
surface of the earth. However, what we actually calculate is
the change in the gravitational potential energy, as the mass
is moved further from the center of the earth, which is a positive change. As it happens, we are generally ignorant of the
total amount of energy in an object, and are only interested
in the change in the amount. If the sign of change is correct,
the sign of the potential energy is unimportant.

1.3.6Power Flows and Signs


The sign convention for energy transfers across system
boundaries varies with engineering discipline and subdiscipline. Engineering sign conventions typically define the
expected, normal, or preferred quantity as positive and the
reverse as negative. For example, thermodynamics uses the
sign convention, where the positive direction of the power
flow depends on the type of energy moving across a system
boundary. For a heat engine, heat flow in is positive and work
flow out is positive. Heat engines cannot be run in reverse.
Turning the drive shaft of an automobile does not produce
fuel. The thermodynamic sign convention of heat in and
work out as positive are the expected flow directions. System dynamics uses a general sign convention that inevitably
conflicts with the sign conventions used in one or more of
the subsystems present in a hybrid system, such as an electric motor. Further, unlike heat engines, ideal linear systems
are reversible. An electric motor can function as generator.
Power sources can both provide and accept power. Oscillations can occur, in which power flows periodically reverse
direction in motors. Although some elements in an energetic
system have a conventional positive direction, most do not,
and, in any case, power can flow either way through all of
the elements.
The generalized sign convention used in system dynamics
varies with the modeling and analysis method. Sign conventions must be defined graphically for the results of the analysis to be interpreted correctly. Once a sign convention is chosen, consistency is essential. The sign convention chosen is
applied to all energy transfers, both across a system boundary and into and out of an individual energetic element. This
text uses the linear graph method and its sign convention,
flow in is positive, flow out is negative. This is the same sign
convention applied to bank accounts. We shall see that sign
conventions within mechanical engineering are essentially
arbitrary. The signs themselves are unimportant, as long as
they are used consistently, so that the accounting of energy
transfer is correct.

1.3.7Power Sources
A common application of system dynamics in machine design is to size the motor needed to drive a system through a
specified motion. To do this properly, we must derive and
solve the differential system equation, which describes the
motion of the machine. We cannot arbitrarily specify both
the magnitude of the force acting to accelerate the mass and
the velocity of its point of application, without violating
F = Ma. We will find a similar constraint in every type of
energetic system we investigate.
Power is provided to energetic systems by power sources.
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,
such as a chemical battery and an electric wall socket. A 12
VDC battery provides a nominal 12 volts at its terminals. An
electric wall socket provides a nominal 120 volts AC or 240
volts AC, depending on location. Note that in both examples,
current is not indicated, only the voltage. The device connected to voltage draws current from the source at a varying
rate, which is dependent on the nature of the specific device.
A battery and a wall socket are voltage sources, because
the voltage is known, not because they only provide voltage.
Voltage by itself is not electric power. DC electric power is
the product of voltage and current.


PDC = vi

(1.35)

AC electric power is calculated differently, because voltage


and current, which vary sinusoidally, peak at different points
in the oscillation, as we shall see in Chap.11.
We can only specify the value of one of the two power
variables associated with any power source. If we attempt
to specify both variables, then we are attempting to dictate
the behavior of the energetic elements in the system. The
behavior of energetic elements is governed by physics, not
by the engineer. System dynamics allows us to predict the
behavior of a dynamic system to a specified input. If we wish
to change the behavior of the system to that input, we must
change the system. The system will draw power from the
source at a rate determined by the magnitude of the power
variable we control and by the nature of the system.
If we measure both power variables during the response
of the system to the power source, we can then calculate the
power drawn from the source by the system. Alternatively,
during machine design, we may pose the hypothetical question, What size (power) force source do we need to move
the point of application of the force with velocity, v(t), if the
force source provides force, F0?
We will illustrate power flow calculation and power source
sizing with the following example. Again, we cannot dictate both power variables of a power source without rewriting
physics. Figure1.18 shows a hypothetical case, in which we

1.3 Energy, Mechanical Power, and Coenergy


Fig. 1.18a Desired velocity v(t)
of the point of application of the
force source. b Hypothesized
force F(t) needed to yield v(t)

17

a
2.0

v(t)
m
___
sec

F(t)
kN

1.5
1.0
0.5
0

Fig. 1.19a Segments for velocity function, v(t). b Segments of


input force function, F(t)

20
15

t, sec

0.5

II

I
0

estimate the force needed to impose a specified velocity at


the point of application of the force, and then calculate the
power the source must provide and the energy required.
Mechanical power is P = F v, Eq.1.17. Although it is not
stated that the force and velocity of the point of application,
Fig.1.18, are colinear, we will assume that they are, and add
that constraint, when we present the results. When the force
and velocity of application of the force are colinear, we can
dispense with vector notation, and express mechanical power
as the product of the scalar variables, P = Fv, Eq.1.31. Notice that the desired velocity v(t) and the hypothetical force
F(t) are discontinuous functions, consisting of straight line
segments. The line segments, v(t) and F(t), have the same durations. In other words, the discontinuities of the two functions
occur at the same times, Fig.1.19. Physically, the changes in
velocity are responses to the changes in force. Notice also that
the input force, F(t), consists of horizontal segments of varying widths, corresponding to constant force magnitudes of
varying durations.
If the functions for both F(t) and v(t) are known, multiplication can be performed to yield the power provided by
the source, P = Fv. We shall see in Chap.3 that discontinuous functions are constructed by summing scaled and timeshifted Heaviside unit step functions.
For the time being, we will evaluate the product graphically, segment by segment, noting that the result of multiplying a velocity segment by a constant force is a segment of
the power function with the same shape as the velocity plot,
except when the product is inverted by the negative force
between three and foursec.

III IV
2

t, sec

t, sec

III

15

F(t)
kN

1.0

20

2.0
1.5

-10

5
-5

a
v(t)
m
___
sec

10

10
0
-5

5
0

IV

II
2

t, sec

-10

The energy provided by the power source is the integral


of power over time.
t

(1.36)
E = P dt E = F (t ) v (t )dt
Unlike the power function, P (t ), the energy function, E (t ),
will be a continuous plot. Power, being the flow rate of energy, can change magnitude or sign instantaneously. The
amount of stored energy cannot change instantaneously, because it is a flow quantity, which accumulates in storage. The
amount of any flow quantity stored in a system or an element
cannot change instantaneously, without an infinite flow rate.
Integration of the power plot can be performed analytically or graphically. You most likely have performed graphical integration when creating shear and bending moment
diagrams. Recall that rectangular areas integrate to ramps or
triangles, while ramps or triangles integrate to parabolas.

adt = at + C

and

at dt = 2 at

+ C2

The constants of integration represent the amount of energy


stored in the system at the beginning of the time interval. If
we assume that the system is de-energized before the input,
F(t), acts on the system, we can neglect the constants of integration.
Graphical integration is performed by calculating the area
of each segment of the power plot, summing those areas to

1 Introduction to System Dynamics

18

40

25

35

20

30

(t) 15
kW sec

25
20

10

III

(t) 15
kW 10

5
0
-5

IV

II
1

4
3

t, sec

-15

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, 40kW. One horsepower equals 746W. Converting from kW to hp, 40 kW

1 hp

746 W

= 53.6 hp

(t)
kWsec

3
4

Shape
Triangle
Rectangle
Rectangle+
Triangle
Rectangle
Triangle

Area
kW sec
2.5
2.5
15

Running
total
2.5
5
20

Integrated
shape
Parabola
Triangle
Parabola

5
10

25
15

Triangle
Parabola

Plot the end points of each segment and connect the dots
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,
Fig.1.21.
To calculate the power provided by the source analytically, we must first create the power function, P
(t ), which
we will then integrate to yield energy, Fig.1.22. P (t ) is the
sum of five separate functions, corresponding to the five
time intervals of Fig.1.20.

20

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 segment integrated. Graphical integration is simple but has the
disadvantage that it only yields values at the ends of the time
segments. Create a table similar to the following:

2.5
3

t, sec

30

-20

IV
V

Fig. 1.21 Plot of energy, E(t), provided by the force source, F(t), in
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,
1 joule = 1 N m = 1W sec

-10

Interval Time-end
points
I
0
1
II
1
2
2.5
III
2

III IV

II

t, sec

Fig. 1.22 Plot of the energy function E(t), produced by analytical integration of the power function P(t)

kW
t
sec

P (t ) = 5

t =1

t =0

+ ( 2.5 kW ) t =1

t=2

kW

+ 20 kW + 40
(t 2 sec)

sec

t = 2.5

t=2

+ (10 kW ) t = 2.5

kW

+ 20 kW + 20
(t 3 sec)

sec

t =3

t=4

t =3

Notice the time shifts (t2) and (t3) created by subtracting the time of the beginning of the interval from the
time variable t. We will introduce time shifts of input and
response functions in Chap.3. Briefly, subtracting the beginning time of an interval from the time variable t shifts time
t=0 to the beginning of the interval, creating a local time
for just that integral.

1.4 Network Representation of Energy Flow

19

E (t ) = P (t ) dt
0

E (t ) = 5
0

kW
t dt + 2.5 kW dt
sec
1
+

2.5

Branch
Node

kW

20 kW + 40 sec (t 2 sec)dt
2

kW
+ 10 kW dt + 20 kW + 20
(t 3 sec)dt
sec
2.5
3
The integrals with time shifts require special handling. We
must define a variable equal to the shifted time. We define
1 t 2 sec for the integral, thus,
2.5

20 kW + 40

Fig. 1.23 Networks transfer a quantity between branches connected


at nodes

The energy function is now:


4

E (t ) = P (t ) dt
0

kW
(t 2 sec) dt
sec

E (t ) = 5
0

d 1 = d (t 2sec ) d 1 = dt d ( 2sec )

d 1 = dt

Substituting 1 into the integral and expressing the limits of


integration in terms of 1 yields

kW
20 kW+40
(t 2 sec) dt =
sec

2.5

0.5

kW
20 kW+40
1 d1
sec

kW

20 kW+20 sec (t 3 sec) dt


3

kW

20 kW+40 sec

d1

kW
2 d 2
sec

The integration is performed piece-wise, evaluating each interval separately.

E (t ) =

5 kW 2
t + 2.5 kW t
2 sec 0

2
1

0.5

40 kW 2
+ 20 kW 1 +
1

2 sec 0

3
20 kW 2

+ 10 kW (t 2.5) 2.5 + 20 kW 2 +
2

2 sec

Likewise, we define 2 t 3 sec for the integral, thus


4

0.5

+ 10 kW dt + 20 kW+20

Differentiating 1:

2.5

kW
t dt + 2.5 kW dt +
sec
1

Back-substitute for 1 and 2 and express the limits of those


intervals, in terms of time t:

E (t ) =
2.5

3
5 kW 2
40 kW
20 kW

2
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
2 sec 0
2 sec
2 sec

Differentiating 2:
d 2 = d (t 3 sec ) d 2 = dt
substituting into the integral and expressing the limits of integration in terms of 2:
4

20 kW+20
3

kW
kW
(t 3 sec) dt = 20 kW+20
2 d 2
sec
sec
0

1.4Network Representation of Energy Flow


A network is a system, in which elements interact with
one another by transferring a quantity via defined paths or
branches which join at connections or nodes, Fig.1.23.
Depending on the type of network, either the branches or
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

1 Introduction to System Dynamics

20

billions of devices which produce or process information. Information 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 system without loss, unless there are leaks, but is also stored by
the system. The water in the system may be changed chemically by the system, if the water is softened, or energetically, if the water is heated. Further, the water flow rate and
flow pressure vary dynamically throughout the system, when
the system is in use.
Information networks process information at nodes and
transmit information without change through the branches,
which connect the nodes. The nodes of an information network act on the quantity flowing through the network. The
branches of an information network conduct (transmit) the
information without change. Conversely, the branches of a
plumbing system act on the water by affecting its pressure
and flow rate. Nodes in a plumbing system are locations selected to measure the changes in pressure throughout the system and to apply mass and energy conservation. The nodes of
a plumbing system are modeled as conducting water, without
storing it or changing its pressure. From fluid mechanics, we
know that any redirection of fluid flow imparts a change in
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
in the system. The process of modeling a plumbing system
assigns the physical properties of the pipes and connections
to branches. The flow of water divides at an ideal node without a pressure change. The branches between the nodes possess physical properties, which affect the waters pressure
and flow rate.
The state of energetic systems changes as energy moves
through the system. We will refer to the movement of energy as its flow. In fluid systems, energy flows with mass. In
electrical systems, energy flows with charge. In translational
and rotational mechanical systems made of solid components, energy is the only quantity which flows through the
system. We will extend the circuit representation of fluid and
electrical systems to translational and rotational mechanical
systems by focusing on power, the flow rate of energy, by
using the linear graph method. A linear graph is a network
representation of an energetic system.
The energetic networks will consist of energetic elements, represented by branches, which connect at nodes.
Each branch represents a different, single, lumped, and fundamental energetic property of the system. These idealized
elements do not directly represent the object or system we
are modeling. They represent a single energetic property or
attribute. In reality, the energetic properties of a real system are distributed in space. The branches have no physical dimensions, just the energetic property. If a component

Pump

Fig. 1.24 A water-filled piping system. The system is horizontal. The


pump increases pressure in the direction of the water flow. The pipes
have fluid resistance. Pressure drops in the direction of water flow
through the pipes

of the real system contains two energetic properties such as


kinetic energy storage and strain energy storage, such that
both are significant enough to be included in the energetic
network, then we will include two branches, with one element for each energetic property. Our network model of an
energetic system simplifies the representation of the system,
while retaining its relevant physical attributes. We will return to the question of how this is done, when we investigate
the modeling of the energetic properties of different types of
components.

1.4.1Compatibility and Continuity Equations


Fluid systems are mechanical systems, since they are described in terms of Newtons laws. Because fluids deform
to fill a volume, Newtons laws are expressed in terms of
pressure rather than force. Consider a horizontal set of interconnected pipes with a pump to circulate water through the
system, Fig.1.24.
If the two following facts are true:
Water flows in the direction of a decrease in pressure,
except through the pump which raises the pressure in the
direction of the flow
Unless the plumbing system leaks, the water flowing
through the plumbing system is conserved
then the piping system can be described in general terms as
a network with a potential driven flow of a quantity which
is conserved.
A potential is the physical variable, which lessens or drops
in the direction of the flow which it drives. Pressure acts as
the driving potential in a piping system. The mass of water
is the quantity, which flows through the piping system and is
conserved. Pressure and mass flow rate are different types of
variables. We can write two sets of summations, one in terms
of mass flow rate, and the second in terms of pressure for
this piping system, knowing only how the pump and pipes
are connected, but not knowing details, such as diameter and
length of the pipes, or whether pressure is applied by the

1.4 Network Representation of Energy Flow

m in

Pump

21

Control Volume

m out

Fig. 1.25
A control volume cutting the piping system. Mass
conservation can be written in terms of mass flow rate for any control
volume drawn on the piping system

pump. These two sets of equations must always be satisfied,


whether or not the pump is running.
If we draw a control volume around any part of the system, such that we intersect one or more pipes, Fig.1.25, then
we can write a statement of conservation of mass in terms
of mass flow rate, by differentiating Eq.1.27 with respect
to time:
d massIn d massOut d massStored
(1.37)

=
dt
dt
dt
Equation1.37 states that the mass flow rate of water into
the control volume, minus the mass flow rate of water out,
equals the rate at which the mass of water, stored inside the
control volume, changes.
In order to systematically extract all of the useful equations from the network, we will find it convenient to draw
control volumes, so that they correspond to single nodes
where pipes connect. Because we can write mass conservation equations at individual nodes of the system, these continuity equations are also called node equations. On the
other hand, a pressure measurement at a single location in
the piping system, such as at a node, does not permit us to
write an equation. We must take pressure measurements at
three locations, if we are to write anything but a trivial equation. (A trivial equation is an identity, such as C=C, which
is true but of no value in algebra.) Useful equations either
sum pressure changes around a loop in the network (the sum
must be zero); or equate the sum of pressure changes from
one node to another node, following two different routes or
paths through the network. Accordingly, equations written in
terms of pressure changes between nodes are either loop
equations or path equations. Equations written using the
driving potential of an energetic system are called compatibility equations, since, for the pressure changes between
nodes to sum properly, they must be compatible.
As it happens, there are always two independent sets
of algebraic equations, which can be extracted from any

network, which has a potentially driven flow of a quantity which is conserved. Independent refers to mathematically 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
of variable, either the flow variable or the potential variable.
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
within the accuracy of the measurements.

1.4.2Compatibility Equations
Pressure is a scalar variable, since it acts equally in all directions. Pressure is also a relative measure. The value of pressure at a point has no meaning, until the reference pressure
measured was relative. Machine design usually uses pressure measured relative to atmospheric pressure, called gauge
pressure, but there are occasions when absolute pressure
is used instead, where the reference pressure is a complete
vacuum.
To describe the behavior of an element in a fluid network,
we need the pressure difference across the element between
the nodes at either end. To describe the behavior of a fluid
system, the pressure differences across all of the elements in
the system must add up. Before we write any compatibility
equations, we must indicate in the schematic of the system
the locations of those pressures of interest to us. The pump
raises pressure, as it pushes water into the system. There are
different pressures on either side of the pump. We will always locate a pressure node where the flow divides or converges. Finally, let us say that the sections of pipe, which join
at the top right corner of the diagram, have different diameters and, therefore, different fluid resistances. We will locate
a pressure node at the location where the diameter changes.
Finally, we will add a reference node for atmospheric pressure, Fig.1.26.
The pressure difference between two points (or nodes) in
the piping system, from node 1 to node 2 can be measured directly, by connecting a manometer between those pressures,
but is more frequently calculated by measuring the pressure
at each point relative to atmospheric pressure and then subtracting.


( p1 patm ) ( p2 patm ) = p1 p2 p1,2

(1.38)

Notice that the reference pressure patm drops out, since it appears in the pressure measurements at both nodes. We will
use p1 as a shorthand for p1 patm , since we usually work

1 Introduction to System Dynamics

22

1
Pump

p1,2

1
Pump

patmosphere

Fig. 1.26 Piping system with pressure nodes at either end of pipe segments and on either side of the pump

with gauge pressure. If no reference pressure is specified,


then atmospheric pressure is implied.
Note that although pressure is scalar, a pressure difference
is directional. For the difference to be meaningful, we must
know the direction of the pressure difference, or drop. The
pressure drop is positive, when the pressure of the first node
is greater than the second node (e.g., p1 > p2 in Fig.1.27).
The pressure difference from node 2 to node 1, Fig. 1.28,
equals the opposite of the pressure difference from node 1
to node 2.
p2 patm ( p1 patm ) = p2 p1 p2,1
p2,1 = ( p1 p2 ) = p1,2
(1.39)
The sign inversion, created by reversing the node subscripts
on the pressure difference, creates sign errors which prove
very difficult to find. We shall see that there must be a positive direction defined for the flow through each branch of a
network. We will customarily use the positive direction of
the flow through the branch as the direction of the pressure
drop. If we need to express the pressure drop in the reverse
direction, we will indicate it, by placing a negative sign on
the positive directions pressure drop.

patmosphere

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
the terms in the pressure difference, i.e., p1 p2 p1,2

If we measure the difference in pressure between all of


the nodes, which form a complete loop at any instant in time,
the pressure differences must sum to zero, since we measured our way around a loop back to where we started. If
the sum does not equal zero, then we have a measurement
error. Summations of the drops in the driving potential of a
system around a complete loop are called loop equations,
Eq.1.39.
p1,2 + p 2,3 + p 3,4 + p 4,5 + p 5,1 = 0
(1.40)
Note that we do not need to know the direction of fluid
flow through a network to state that the pressure differences
around a loop sum to zero. The sum of pressure differences
around a loop in one direction (clockwise or counterclockwise) must equal zero, because the pressures at each node
are measured relative to the same reference, atmospheric
pressure. Each pressure appears twice, once with a sign
inversion. Expanding the sum of pressure differences:

( 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

p2,1

23

p1,2

3,4

Pump

patmosphere

1,5

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

p1,2

p2,3

3
p

3,4

1
5,1

3
p

p2,3

Pump

Pump

atmosphere

p4,5
Fig. 1.30 Arrowheads indicate the positive direction of the pressure
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 arrowhead is reversed only forthe pressure drop across the pump. Power
sources, such as pumps, raise the driving potential in the direction of
the flow through them

p1, 2 + p 2,3 + p 3, 4 + p 4,5 + p 5,1 = 0


p1, 2 + p 2,3 + p 3, 4 + p 4,5 + p 5,1 p 5,1 = p 5,1

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
Fig. 1.29 Positive directions of flow through the pump and piping.
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,
Eq.1.39

Pressure drops in the direction of steady fluid flow, except


for the flow through pumps. Pumps energize a fluid system
by raising pressure in the direction in which the pump pushes
the flow. The positive direction for the flow through the pump
is shown in Fig.1.29. However, the pressure drop through
the pump occurs in the opposite direction, from node 1 to
node 5, as shown in Fig.1.30. The alternative to a loop equation is to sum the pressure drops between two nodes along
the different paths between them in a path equation. The
path equation for the pressure drop from node 1 to node 5 is
p1,5 = p1,2 + p 2,3 + p 3,4 + p 4,5
(1.41)
Because reversing the order of the node subscripts inverts
the sign of a pressure difference, we can write the loop equation Eq.1.39 as the path equation Eq.1.40, by subtracting
the pressure p5,1 from both sides of Eq.1.39:

We will now generalize. The driving potentials in fluid systems (pressure), electric circuits (voltage), and thermal systems (temperature) are scalar variables. Summing the differences of the scalar potential variable around a loop in a
system creates a loop equation. Alternatively, equating the
sums of the differences in the scalar potential variable between two nodes along different paths creates a path equation. Both loop and path equations are termed compatibility equations. The term comes from geometric compatibility
mechanical design. The linear graph method will allow us to
represent energetic mechanical systems are networks. When
we do, we shall see that geometric compatibility provides a
set of equations analogous to the compatibility equations of
fluid, electrical, and thermal systems.
The reduction of individual equations which describe independent aspects of an energetic network to a differential
system equation which describes the dynamic response of
the system is largely a process of successive elimination by
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
reduction. If we use dependent equations in our algebraic reduction, we will be prone to derive the true but trivial result
0=0, instead of the differential equation we need. It can be
difficult to spot redundant or dependent equations, particularly when we work with larger systems. Consequently, we

1 Introduction to System Dynamics

24

1
Pump

patmosphere

Holes

Fig. 1.31 The topology of a network are aspects which do not


change, when the network is stretched or deformed, as long as the elements remain connected. One aspect of the topology is the number of
holes in the network

will establish a few simple rules to prevent the inclusion of


redundant or dependent equations in our description of the
physics of a network.
The topology of a network refers to how the branches
are connected. We will place only two requirements on how
branches are connected to ensure that our networks will be
proper energetic networks.
1. Branches have only two ends, each of which must be connected to a node.
2. There are no dead ends. It must be possible to travel
through every branch in the network without reversing
direction.
The first requirement implies that the flow through the system can only divide at nodes. Branches do not divide the
flow through by being wye or tee shaped. In other words,
branches do not branch. Branching occurs at nodes. The

Fig. 1.32 Only two of the three


paths are linearly independent

2
1
Pump

second requirement ensures that the network is a circuit. It


will be more important when we draw network representations of translational and rotational systems.
There are aspects of a networks topology which are
invariant and do not vary or change, unless the network
is changed. The invariants are the number of nodes and
holes in a network, Fig. 1.31. A hole in a network is similar to a hole in a fishing net or window screen. It is a region 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
example network has two holes.
The number of independent compatibility equations that
can be written for a network is less than or equal to the number of holes in the network. Often, only a few of the possible equations are useful, while the remaining are either
redundant or trivial. The example network has two holes
and, therefore, a maximum of two independent compatibility
equations. There are three compatibility equations which can
be written, Fig.1.32. Only two of the three can be linearly
independent.
If the piping system is simplified slightly by removing
node three, Fig.1.33, the number of holes in the network
remains unchanged. However, only one useful compatibility
equation can be written. One of the remaining compatibility
equations is redundant, and the other is trivial.

1.4.3Continuity Equations
Continuity equations are written by applying conservation of mass to the flow into and out of the nodes. We will

4
patmosphere

1
Pump

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


Fig. 1.33 Three possible path
equations are presented, but only
one is useful. The other two are
redundant or trivial

Pump

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

patmosphere

1
Pump

1
Pump

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

4
patmosphere

4
patmosphere

Path 2-4 = 2-3-4

4
patmosphere

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

2
1

1
Pump

4
patmosphere

Path 2-4 = 2-4

1.4 Network Representation of Energy Flow


Fig. 1.34 a Branches need
unique identifications. Normally,
they are named using a parameter of the branch. b Identifying
branches by the nodes they are
connected to does not work,
since parallel branches receive
the same identification. This
network has two branches named
Branch 2,3

25

Branch A

Branch A 2

Branch B

patmosphere

Pump

Branch 3,4

Branch C
patmosphere

Branch E

later find it convenient to model the water as incompressible, and express the equations in terms of volume rather
than mass. Continuity equations are written at nodes and
are independent of the energetic properties of the branches.
Although there are many types of networks in which nodes
have properties, such as transportation and information networks, we have defined a node as a location without physical properties. In the energetic networks we will work with,
the branches represent the different energetic properties of
the system. The nodes exist to connect the branches to one
another. Nodes cannot store anything.
In order to write continuity equations, we must:
1. Identify the branches.
2. Assign a positive direction to flow in each branch.
We need an identity for each branch, in order to identify the
flow through the branch. We cannot identify the flow through
a branch by the nodes at either end, because parallel branches which are connected between the same nodes have different flows through them, Fig.1.34. Normally, we will identify
a branch with the parameter from the energetic equation for
that branch. For example, if the branch represented a spring
with spring constant, K, we would identify the branch with a
K. This is the same convention used to identify elements in
an electrical schematic. For this example, however, we will
simply identify the branches alphabetically.
Defining the positive direction for flow through a branch,
by drawing an arrow pointed in the positive direction, is

mA

mB

Branch D

Branch 2,3

Branch D
Fig. 1.35 a Each branch in the
system is given unique identification. The positive direction for
each branch is indicated on the
schematic. b Mass flow rates
used in continuity equations

Branch 2,3

Pump

patmosphere

Branch 1,2 2

Branch C

Pump

Branch B

mD

mPump Pump
5

mC

patmosphere

mE
termed orienting the branch, Fig. 1.35. The direction we
choose as positive for the flow in a branch is arbitrary for
most elements. The mathematics will yield the correct flow
direction by inverting the sign of a flow, if necessary; similar
to how arbitrarily assigned directions of forces in the members of a truss are corrected by the truss analysis. Power
sources are an exception. The positive direction of the flow
is not arbitrary. The positive direction of the flow is in the
direction the driving potential increases. Motors and power
transmission elements are the only other energetic elements
that have defined positive flow directions, as we shall see in
Chap.6.
Recall that we have adopted the sign convention that flow
into a node is positive, and flow out of a node is negative.
Just as compatibility equations can be written as either loop
or path equations, continuity equations can be written in two
ways: (1) equating the sum of the flows into and out of the
node with zero; or (2) equating the sum of the flows into the
node, with the sum of the flows out. There are five nodes in
this system. Therefore, we can write five continuity equations. We shall see that only four of the five continuity equations are independent. Any one of the five continuity equations can be derived from the remaining four. Writing the
mass conservations equations in the form, flow in= flow
out, for all five nodes, starting with the flow out of the pump,
leads to the following:

1 Introduction to System Dynamics

26
Engineering
Objective

Design
Identify a Question
Design
Decision

Create an
Engineering
Model

Engineering
Numerical
Design
Model
Analyze Result
Interpret Decision
Model
Result

Fig. 1.36 Block diagram of an


engineering design decision

Focus of Undergraduate
Engineering Education

m Pump = m A
m A = m B + m C
m B = m D
m C + m D = m E
m E = m Pump
We can derive any one of these continuity equations from
the remaining four. We will demonstrate their dependence
by deriving m E = m Pump from the first four continuity equations, using elimination by substitution, and starting with the
continuity equation, m Pump = m A .
m Pump = m A m Pump = m B + m C
m Pump = m D + m C m Pump = m E
We do not want to include dependent equations in the set of
equations we will develop to describe energetic networks,
because it can greatly increase the effort of reducing the set
to an equation which describes the dynamics of the system.
Consequently, our rule will be to always write one fewer
continuity equation than there are nodes in the network. It
will be convenient to adopt the habit of omitting the continuity equation for the lowest potential node.

1.4.4Summary of Compatibility and Continuity


Equations
Compatibility equations are mathematical statements that
the differences between a scalar variable, such as pressure,
voltage, and temperature, measured between different locations in an energetic system must sum to zero, when we add
the differences around a loop and return to our starting point.
Likewise, if we measure the differences between the scalar
variables at different locations along two different paths between the same two starting and ending locations, the two
sums must be equal.
Continuity equations in fluid, electrical, and thermal
systems are statements of mass conservation, charge conservation, and energy conservation, respectively. Nodes are
the locations, where energetic elements connect in energetic networks. Nodes have no energetic properties. Since
they cannot store mass, charge or energy, all mass, charge,

and energy which flows into a node must immediately


flow out.
The number of useful compatibility equations which can
be extracted from an energetic network is unknown, but
its maximum value is limited to the number of holes in the
network. The number of independent continuity equations
which can be extracted from an energetic network is one
fewer than the number of nodes in the network.

1.5Overview of Engineering Modeling


and Analyses
The focus of a modern engineers education is on developing
the mathematical skills and understanding needed to derive
and evaluate equations based on physical theory. There are
very good reasons to structure an undergraduate engineering
curriculum this way, but the student is often unaware that the
derivation and evaluation of an equation from an existing
engineering model are only intermediate steps in the process
of making an engineering design decision, Fig.1.36.
We use the term, analysis, loosely to refer to engineering computation. However, engineering computation is not
analysis. Analysis is the opposite of synthesis, which
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
model. Textbook problems do not teach engineering analysis
since the system has already been taken apart and modeled.
An unfortunate but common result of the introductory
courses of an engineering education is that engineering students develop unrealistic ideas regarding:
Why engineering calculations are performed,
The accuracy of the information used to perform an engineering calculation,
The uniqueness of the model the calculations are based
on,
The confidence which can be placed in the results of engineering calculations, and
The manner in which they should approach a novel or
unfamiliar problem.
These engineering student myths are eventually dispelled
by experience after the engineering graduate begins his or
her professional practice, but it is far better to spare the student the pain of learning through error or embarrassment and

1.5 Overview of Engineering Modeling and Analyses

expose the myths now. In a nut shell, the study of engineering is quite different from the practice of engineering.
Engineering analyses are tools used by engineers to aid
design decisions. The results of an engineering computation
allow an engineer to make a rational decision between alternatives, or to validate a decision made using engineering
judgment. Engineering calculations are performed to provide a basis for engineering design decisions. Any type of
design, whether engineering design, graphic design, or landscape design, is a sequence of decisions. Design decisions
are constrained by economic, physical, and practical factors.
For example, a graphic designer may wish to use a size and
quality of paper for a project but cannot without exceeding
the projects budget; a floral designer may wish to use tropical plants for an installation but cannot, because they would
not withstand a northern climate; and a mechanical engineer
may wish to use a manufacturing process that the client does
not possess.
Additional constraints are imposed by design practice.
Designers and engineers of today use design elements which
may data back years, centuries, or millennia. The design
question, What diameter should the shaft be? was asked by
ancient engineers. The shafts we design today usually look
like ancient shafts, since they typically have a circular or,
if not circular, an axially symmetric cross-section such as a
square or hexagon. The modern engineer has more materials and more methods of shaping or processing materials to
choose from than his or her ancient predecessors. Also, modern engineers can use physical theory expressed in mathematics to describe materials, and how loads applied to a shaft
of a given geometry affect the materials to aid their design
decisions, which the ancients did not possess. However, the
objective remains unchanged: make justifiable decisions as
to the material and dimensions of the shaft, so that it will
serve its purpose at an acceptable cost for a reasonable life
time.
Although ancient engineers computed some quantities,
such as volumes and weights, their lack of physical theory and modern mathematics forced their design decisions
to be based on empiricism (i.e., use the same diameter as
the last shaft with similar loading which worked). Modern
engineers still use empiricism. Any change in a successful
design requires rational justification. In some cases, empiricism is used because there is a substantial lag between the
development of a new material or process and development
of mathematical theory to describe it. In other cases, empiricism is used because the data needed to apply mathematical
theory are unavailable, or so scattered or variable over space
or time, that the mathematical theory yields contradictory results. However, reliance on empiricism can be expensive and
limiting. It can be expensive, because the only way to optimize an empirical design, i.e., to create the least expensive
design which serves its purpose, is by comparison between

27

successful designs and failures. Most often, the information


regarding failures is not available, either because it does not
exist, since the failure loading was unknown, or it is proprietary, meaning it is a trade secret. The limiting aspect
of empiricism is obvious. Empiricism cannot be applied to
anything novel, since there is no experience base to build on.

1.5.1Engineering Modeling and Analysis


Process
The common aspects of the various engineering analyses
you have studied was present in Sect.1.0.6 as a generalized
process of engineering modeling and analysis divided into
four steps:
Step 1. Draw a picture(s) of what you have chosen to analyze, to aid your engineering decision making. This
is your model. Annotate the picture(s) to define all of
the variables and parameters in your model.
Step 2. Write mathematical statements of relevant physical
truths.
Step 3. 
Reduce the mathematical statements, eliminating
all unknown variables except the input and output
variables, and solve the equation (or system of equations) for the output variable(s).
Step 4. Evaluate the solution. Iterate if necessary.
We will now elaborate and describe how this process will be
applied to model and analyze dynamic systems.
Step 1. Define the model by drawing a picture Graphical
representation is the first step in engineering modeling. If you
look back at your courses to date, every engineering analysis
started with a picture. In statics problems, your instructors
stressed the need to draw pictures (i.e., free body diagrams),
by penalizing you points, even if you reached the correct
solution. Why such a hard line? Engineering graphics are
not just an embellishment to make engineering computations
more attractive. Engineering graphics establish the model
and define the parameters and variables. Modeling is the
most important step in engineering analysis. A very approximate solution to a reasonable model is helpful, whereas a
very precise solution to the wrong model is dangerous.
The relevant attributes needed to model the dynamic
characteristics of a system are these: how is energy supplied,
transformed, stored, and dissipated. Most dynamic models
use ideal, linear element equations which assume that the
way energy is transformed, stored, or dissipated can be described by a constant parameter. This is never really true,
but it is generally reasonably accurate. We typically must
include more ideal elements in our model than we see in the
physical object, since our model represents each attribute individually; whereas, in reality, all of the attributes are combined in the physical object.

28

Modeling is the most interesting step of engineering


analysis because it requires the most judgment. In systems
dynamics, we will use lumped parameter models. You have,
in fact, already drawn many lumped parameter models. Free
body diagrams are examples of lumped parameter models,
since physical properties that are actually distributed, such
as mass, are represented by a single parameter. Electrical
schematics are also lumped parameter models. For example,
electrical resistance is present in all conductors at room temperature. However, in electrical schematics, resistance is
only indicated for discrete components named resistors in
order to keep the lumped parameter model simple enough
to reduce it mathematically. The wires connecting components have resistance also, but we only include significant
amounts of resistance in the model.
How does one judge significance? Sometimes, it is obvious because there can be differences by orders of magnitude between the amounts of a given physical property in
two elements of a system. At other times, it may not be at
all obvious, and may require iterating the model. We will
use a rule of thumb for the first model, such that an element
which stores or dissipates less than 10% as much energy/
power as the next smallest similar element in the system can
be neglected.
Once we have established a model, the reduction and solution is algorithmic, and in engineering practice is done by
a computer. Consequently, modeling is the key step in the
analysis. If the model (whether it is a free body diagram,
electrical schematic, control volume, or lumped parameter
model) is wrong, the entire analysis is at best useless, regardless of the sophistication of the reduction and solution.
If the model is correct, one can often make the necessary
engineering decisions simply by looking at the picture. In
more complex situations, you must proceed from a graphical
model to a mathematical model.
The graphical model (picture) must be annotated with all
relevant parameters and variables before proceeding to the
second step of writing equations. If in Step 2 you find you
have forgotten a parameter or a variable, do not neglect to
add it to the picture. Failure to show all parameters and variables graphically results in errors, particularly sign errors,
and miscommunication.
Step 2. Write mathematical statements of relevant physical truthsMany engineering students mistakenly believe
that engineers faced with an unfamiliar problem can think as
clearly and efficiently as analyses presented in engineering
text books. If you have ever asked yourself, How did the
author know to introduce that equation at this point in the
analysis? it may well be that the author didnt the first time
he or she solved the problem. Engineering texts are revised
through many drafts to be as clear and as efficient as possible.

1 Introduction to System Dynamics

Engineers do not solve unfamiliar problems with either


the efficiency or the structure of a textbook analysis. The
first step an engineer takes is to ask him or herself, What
do I know about this model? This step typically involves
rough sketches and lists of unrelated equations, some general
principles, others specific to problem at hand. It is only after
producing these problem-solving resources that the engineer
attempts to impose some order and logic. It is seldom successful, or even helpful, to begin a mathematical derivation
immediately.
The general method of engineering analysis is to separate
Step 2, the process of writing equations that represent the
various physical truths of the system, from Step 3, the reduction of those equations to the inputoutput relationship.
Separating these processes into two distinct steps helps to
both identify all of the relevant physical truths, and ease the
mathematical reduction, as will be demonstrated in the examples that follow.
What we is need a complete set of independent equations
which describe the system. In most cases, the statements of
physical truth in an engineering problem can be grouped into
five independent categories, the order of which is irrelevant.
Although the flow or creation of entropy, the fifth category
below, is important in reality, our models will not include its
effects, other than viewing energy dissipation in the form of
heat as irreversible.
Independent Physical Truths:
1. Compatibility of displacements, velocities, voltages temperatures, and pressures. (Loop or Path Equations)
2. Continuity of current, Conservation of mass, or Equilibrium of forces and torques. (Node Equations)
3. Material or Elemental Model. (Constitutive Equations)
4. Energy Conservation and Power.
5. Entropy Creation and Flow.
1. Compatibility: A fundamental physical truth is that, if you
go around a loop, you get back to where you started. If you
measure your progress around a loop in steps, the steps must
be compatible, in the sense that the sum of the steps must be
equal to the whole. Fluid systems must have pressure drops
across components that add to zero, when you go around a
loop. Electrical systems must have voltage (potential) drops
across components that add to zero around loops. Thermal
systems must have temperature drops that add to zero around
loops. Mechanical systems must have geometric compatibility of displacement, velocity, and acceleration. The deformations of the individual parts must be compatible, in that they
must add up to the deformation of the whole object.
2. Continuity, Conservation, or Equilibrium: Another category of fundamental physical truth is that matter and energy

1.5 Overview of Engineering Modeling and Analyses

can either stay put or go somewhere, but they cannot just


disappear. They are conserved. The conservation principle
can be applied by measuring flow rates. In the special case
of an element that cannot store matter or energy, what flows
in must flow out. In fluid systems, if an incompressible fluid
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
must flow out the other. In thermal systems, if heat flows in
one end of a metal bar, it must flow out the other.
Statements of equilibrium in mechanical systems are
analogous to continuity since they involve quantities, forces
or torques, which sum to zero at a point or on a body. We
will use dynamic equilibrium which includes as a force
the net force which accelerates mass. The net force acting to
accelerate a mass is the opposite of the DAlemberts inertial
force. The inertial force is the force that pushes back, when
a mass is accelerated. When the net force acting to accelerate
a mass, which we will call FM, is included in an equilibrium
statement, the force summation equals zero. This dynamic
equilibrium is a method of accounting for all forces acting at
any point in a system.
3. Material or Elemental Model: Compatibility and Continuity equations are summations which equal zero. There is
only one type of variable in each summation. Force summations have only forces. Pressure summations have only pressures. The material or elemental model equations relate two
different variables by means of a parameter. The equation
which describes a spring element is familiar F = Kx, where
F is the force in the spring, K is the spring constant (also
called the spring rate), and x is the deformation of the spring.
This equation relates two different types of variables, force
and displacement. The spring constant is the parameter. The
equation for a resistor is another such elemental equation
v = iR, which relates two different types of variables, i.e.,
voltage and current. Resistance is the parameter. The concept of a parameter carrying units in order to relate dissimilar
variables is attributed to Fourier. Previously mathematical
statements were proportionalities, not equations.
We will work with elemental equations expressed in
terms of power variables of the system. The product of voltage times current is electric power. Hence, voltage and current are the power variables for an electrical system. The
familiar equations for electrical system elemental equations
are written in terms of electrical power variables. However,
the familiar equations for mechanical systems are not.
In system dynamics, we view masses and springs as mechanical energy storage elements. We will use unfamiliar
forms of the elemental equations, F = Kx and F = Ma, in
order to express the equations in terms of the power variables
of mechanical systems. We will also restrict each element to
motion or deformation in one dimension. If deformations in
two or more dimensions are coupled, they will be represent-

29

ed as if they occurred in one dimension. The restriction to


one dimension allows us to work with scalar equations. The
elemental equation for a mass is Eq.1.29:
F = Ma
Rewriting acceleration in terms of the power variable velocity yields the differential relationship between the power
variables of a mechanical system, F and v, per Eq.1.30:
F=M

dv
dt

Note that the time derivative is on the power variable velocity which determines the kinetic energy stored in the mass.
It takes time to transfer energy into or out of a mass, hence
the derivative on the energy storage variable for the mass,
velocity.
Hookes law, Eq.1.6, is the elemental equation for a
spring,
F = Kx
Rewriting displacement in terms of the power variable, velocity, requires differentiating both sides of the equation with
respect to time:
dF
dx
=K
dt
dt
dx
= v, yielding a differential equadt
tion for elastic strain energy storage:

and then substituting for

dF
= Kv
dt

(1.42)

Writing the equations for mechanical energy storage elements in terms of the mechanical power variables, F and v,
allows us to develop mathematical analogies between different types of physical systems, because the equations have the
same form. For example, energy is stored in an electric circuit in either the electric field of a capacitor or the magnetic
field of an inductor, Eqs.1.43 and 1.44,
dv
(1.43)
i=C
dt
di
(1.44)
v=L
dt
where C is the capacitance in farads, and L is the inductance
in henrys. Notice that the equations for the electrical energy
storage elements are also differential equations.

1 Introduction to System Dynamics

30

Compare the elemental equations for a capacitor and an


inductor with those of a mass and a spring:
dv
dt
1 dF
Spring: v =
K dt
Mass: F = M

dv
dt
di
Inductor: v = L
dt

Capacitor: i = C

Even though mechanical and electrical systems are very different, the equations which describe them are not. Energy
storage elements have a differential relationship between
the power variables of the system. The time derivative is on
the variable, which determines the amount of energy stored
in the element. The coefficient carries the units needed to
equate two physically different variables and scales them to
the set of units, SI or US customary, being used.
The linear graph method uses these analogies between
different types of physical systems, which will allow us to
create a single model to represent devices which contain different subsystems, such as an electric motor which has both
electrical and mechanical elements.
4. Energy Conservation and Power: We can write mathematical statements which express the amount of energy
stored in an element or the rate at which power is dissipated
by an element. Energy storage equations always contain
an elemental parameter, and all but thermal capacitance
and gravitational potential energy have a squared term. For
example, the amount of kinetic energy stored in a mass is
Eq.1.34:

EM =

1
Mv 2
2

The amount of energy stored in a spring can be expressed in


terms of either the applied force or the resulting deformation,
Eqs.1.11 and 1.12:

EK =

We also see similarity in the mathematical form for these


physically different energy storage elements, as seen in their
element equations.

1 2 F2
Kx =
2
2K

The amount of energy stored in the electric field of a capacitor is the following
1
EC = Cv 2
(1.45)
2
where C is electrical capacitance. The amount of energy
stored in the magnetic field of an inductor is the following
1
E L = LiL2
(1.46)
2

1
Mv 2
2
1 1 2
Spring: E K =
F
2K
Mass: E M =

1 2
Cv
2
1
Inductor: E L = LiL2
2

Capacitor: EC =

The linear graph method uses physical analogies to classify


power variables. There are two possible cases. The power
variables are the same at either end of an energetic element,
such as the flow through a pipe, or they are different at either end of an element, such as the pressure drop across a
pipe. When a power variable is the same at both ends of an
element, it either acts or flows through the element. It is a
through variable. When the power variable can be different at the two ends of an element, or changes across an
element, it is an across variable. An equally useful set
of analogies groups variables by those which represent an
effort, and those which represent either flow or motion.
This set is used by a different modeling method known as
Bond Graphs.
Analogies are valuable, because they help leverage our
knowledge, so that we may use what we understand, in order
to gain insight into what we do not understand. Analogies
also simplify mathematical analysis. Analogies between
different energetic systems have limits, since they compare
dissimilar systems. Thermal systems are energetic systems
which have a driving potential or across variable, temperature, and a through variable, heat flow rate. Although we will
be able to represent thermal system using the linear graph
method, some analogies and calculations will not be applicable. In all systems except thermal systems, the product of
the through variable of an element and the difference of the
across variable at either end is the power flow into or out of
the element. In thermal systems, power is the through variable, heat flow rate. Thermal systems have only one type of
energy storage, thermal capacitance, energy stored in mass
as heat. The amount of thermal energy stored in a mass appears below,


ET = C p MT

(1.47)

in which Cp is heat capacity. Note that temperature T is not


squared. We shall see in Chap.5, that the energetic equations
for thermal systems do not follow the same mathematical
patterns as those for mechanical, fluid, and electrical systems.
Step 3. Reduce the mathematical statements, eliminating
all unknown variables except the input and output variables,

1.5 Overview of Engineering Modeling and Analyses

and solve the equation (or system of equations) for the output variable(s).
By separating Steps Two and Three, the derivation of the
system equation (i.e., the differential equation that represents the dynamics of a system) becomes a straightforward,
although sometimes laborious, process of reducing the mathematical 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 differentiating with respect to time, but never integrating. It is
literally possible to start the reduction by randomly choosing
any compatibility, continuity, or elemental equation, with the
exception of trivial equations of the form, x=x. The energy
equations are not used to derive the system equation. They
are used to provide the initial conditions to solve the differential equation, as we shall see.
Note the use of the verb, solve, not the verb, integrate. Our purpose is to derive a mathematical model that
allows us to predict how a physical system will change with
time. Our purpose is not to investigate various techniques of
integration. We will be able to represent high-order dynamic
systems as the superposition of first and second-order responses. Consequently, we will only need to solve first- and
second-order ordinary differential equations with constant
coefficients. The verb, solve, means to find the solution.
Because we are dealing with physical systems, we will already know the forms of the possible solutions. In the case of
first-order differential equations, there are only two possible
responses, and solving will be simply identifying which
response is reasonable, and calculating the appropriate coefficients to use in the standard form. Second-order dynamic
systems have a much broader range of possible responses.
Their solution is, consequently, more involved and will require the use of complex variables. However, we will again
know the standard form with which we are working, and the
solution will involve finding undetermined coefficients or
manipulating an equation to match a known Laplace transformation.
It is essential to explore the analytical solution of second-order systems to understand the response of dynamic
systems. However, once that understanding is gained, it is
much more productive to use numerical solutions. We will
use computational software, either Mathcad or MATLAB,
to solve and plot the response of our system equations to a
variety of inputs. Mathcad or MATLAB will also allow us to
model higher-order dynamic systems as a set of simultaneous differential equations.
Step 4. Interpret the results of the calculation Every engineering calculation requires interpretation. The first question
the engineer must ask him or herself is Do I believe these
results? Do they seem reasonable within your experience?

31
Rigid end cap

Bar 1, Area A 1 , Modulus E 1

Bar 2, Area A 2 , Modulus E2

L
Fig. 1.37 A mechanical system consisting of two dissimilar elastic
bars of unloaded length L connected to a rigid end cap. Force F acts
on the end cap

Engineering students have little experience so, at first, they


should only expect themselves to identify and reject extreme
or absurd results. With time, however, the calculations performed as an engineering student serve as the beginning of
professional experience, if the student takes the time to think
about the result instead of hurrying onto the next problem.

1.5.2Engineering Modeling and Analysis


Examples
The general process of engineering modeling and analysis is illustrated with two examples. The first example is
from introductory engineering mechanics, and the second is
from electric circuits. The systems are clearly very different. However, the general process of modeling and analysis is the same. The only difference you may see between
how these examples are worked, and textbook examples
you have studied in the past, is the distinct separation of
the steps of: (1) annotating the engineering graphic of the
physical model; (2) extracting mathematical statements of
physical truths; and (3) reducing the set of equations to an
inputoutput relationship. Your first impression may be
that this general method is inefficient, which is true in most
cases. There is nothing more efficient than having a formula
to yield a needed result. However, working a textbook problem designed to illustrate the use of a formula and addressing a novel or unfamiliar design question are very different
experiences. It is in the latter case, that a general method is
the most efficient.

1.5.2.1Example 1: Mechanical Design Analysis


Design Question. Two dissimilar metals bars with Youngs
moduli E1 and E2 have the same length L but different crosssectional areas A1 and A2, as shown in Fig.1.37. A force F is
applied by an ideally rigid member such that the bars elongate without rotation. Find the stress in bar one.

1 Introduction to System Dynamics

32

F1
F2

x1
x2

x
F1
F2

of physical truths helps organize both the thought and the


presentation of ones work.
1. Compatibility: Geometric compatibility tells us that if the
bars elongate without rotation, then the displacements x1 and
x2 are equal.

L
Fig. 1.38 Free body diagrams of the forces acting on the bars and on
the rigid end cap, as well as displacements of the ends of the two bars
and the rigid end cap

Step 1. Draw and annotate a picture (or pictures) The


geometry of the structure is shown Fig.1.37. We need to
draw free body diagrams of the bars and the end cap, in order
to define the axial forces in each bar, establish an equilibrium relationship between the external force F applied and
the forces applied by the bars to the end cap, and to establish
the relationship between displacement of the end cap and the
displacements of the two bars. We would also draw free body
diagrams of the individual bars, if we needed to determine
the reaction forces, Fig.1.38.
It is essential to recognize that engineering models are
conveyed by the annotated engineering drawings we begin an
analysis with. Figure1.37 implies two important constraints,
which greatly simplifies the analysis. The constraints are not
stated. They are implied by omission.
1. All of the vertical dimensions needed to calculate the
moment on the end cap due to the applied forces are omitted. If this were a real system, the end cap could rotate,
unless external force F was applied at a point, such that
the moments created by the bar forces F1 and F2 were
equal and opposite. Moment equilibrium for the end cap
would be provided by the rigid connections between the
bars and end cap and the bars and the support. We will
accept the constraint of purely horizontal displacement.
2. The external force F is not defined. The mechanical system is an elastic structure which stores strain energy. We
concluded in Sect.1.2, that it is impossible to instantaneously impose a finite force on an elastic object, without providing infinite power. Consequently, there is an
implied dynamic constraint that, either the external force
F(t) is applied gradually, i.e., ramps up from zero, to limit
the power needed or the analysis is only valid in steadystate, after all of the transient fluctuations in force and
displacement have ended.
Step 2. Write mathematical statements of physical
truth A systematic approach to identifying physical truths
which can be expressed as mathematical statements makes
this step much easier. Typically, when students are stuck on a
problem, it is because they have omitted one or more physical truths. A mental checklist of general categories or types

x1 = x2 = x
2. Continuity, Conservation, or Equilibrium: The free
body diagram allows us to write an equilibrium equation.
F = F1 + F2
3. Material or Elemental Models: We can relate stress and
strain in both materials.

1 = E11

2 = E2 2

We can also write equations that relate displacements to


strains and forces to stresses
x1
L
F1
1 =
A1

1 =

x2
L
F
2 = 2
A2

2 =

and write equations for the spring constants (or spring rates)
of the bars.
K1 =

F1
x1

K2 =

F2
x2

4. Energy and Power: The amount of strain energy stored


each bar is

E1 =

1
K1 x12
2

E2 =

1
K 2 x22
2

The total amount of energy stored in the system is the sum of


the strain energy in both bars.

E = E1 + E 2
Because the system is elastic, the total energy in the system
can be written in terms of the applied force F and the displacement x:

E=

1
Fx
2

Note that the elastic energy E Fx because F is not constant. F increases linearly with displacement.

1.5 Overview of Engineering Modeling and Analyses

Step 3. Reduction and Solution Having written mathematical statements for all physical truths we can from the engineering model, the remaining effort is one of pure algebra.
We will reduce the set of equations we created, using elimination by substitution. It is helpful to state the input variable
acting on the system and the variable desired as the result
or output of the analysis. The input variable acting on the
system is the external force F. The desired output variable is
the stress in bar number one, 1. Our objective is to produce
an equation with only the variables, F and 1. Parameters can
remain, since, in fact, they are necessary, but no variables,
other than the input F and the output 1, are permitted in the
final equation.
Start the reduction by choosing any equation in the list,
other than a trivial equation. Trivial equations are algebraic
dead ends which lead to the result, 0=0. There is no right
equation to start with. If there is an equation with the input
and output variable in it, that would be a good choice. Otherwise, start with an equation with either the input or output
variable. We will reduce the equation list twice, starting with
two different equations to demonstrate that it does not matter
which non-trivial equation we start with.
Reduction One: Input F, Output 1. Lets start with this
equation:
F
1 = 1
A1
The only variables allowed in our result are the input F and
the output 1. A1 is a parameter. It stays in the equation. F1
is an unwanted variable. We will substitute to eliminate it.
Look in list of equations for one with F1 in it and use it. It
doesnt matter which equation with F1 we use. We can use
the same equation twice if needed, but it we immediately use
the same equation, we undo the previous substitution. There
are three equations in the list which contain F1:
F = F1 + F2

F
1 = 1
A1

33

tion in your list with F2 in it, and use it to eliminate F2. We


will use

2 =
substituting

1 =

F = F1 + F2 F1 = F F2
which yields
F F2
1 =
A1
We have introduced the input F and the bar force F2. The
input F stays in the equation. F2 is unwanted. Find an equa-

F 2 A2
A1

A2 is a parameter. It remains in the equation. We now substitute to eliminate 2 using

2 = E2 2
which yields
F E 2 2 A2

1 =

A1

We have introduced a parameter, E2, and a variable, 2 . Eliminate 2 by substitution. Use the definition of strain 2

2 =

x2
L

to yield

1 =

F E2
A1

x2
A
L 2

We have again introduced a parameter, L, and a variable, x2.


Eliminate x2 by substitution. Use of
x1 = x2
yields

F
K1 = 1
x1

We will use

F2
F2 = 2 A2
A2

1 =

F E2
A1

x1
A
L 2

Eliminate the displacement x1. Using the definition of strain


1
x
1 = 1 x1 = 1 L
L
yields:

1 =

F E2

1 L

A1

A2

F E 2 1 A2
A1

Eliminate the strain 1 : Use Hookes law

1 Introduction to System Dynamics

34

1 = E11 1 =

E1

to yield
F E2

1 =

E1

A2

A1

We now have an expression that has only the input variable


F, the output variable 1 , and parameters. We are done substituting. We are not yet in a useful standard form. We rearrange this equation to isolate the output variable on one side.
This process of rearranging is known as solving for a variable, since it yields an equation for that variable.
E2

E1

1 =
1 +

E 2 A2

1 =

E1 A1

1 =

F E 2 A2

1
A1 E1 A1

E 2 A2
F
F
1 +
1 =
A1
E
A
A

1
1
1

We wish to present our final result as a proper ratio, not as a


ratio of ratios. If we divide both sides to clear a factor from
one side, we will then have ratios to clear later. Rather than
divide, we will save effort and reduce errors, if we place the
left side term over a common denominator, so that we can
clear it from the output variable 1 by multiplying both sides
by its inverse.
E 1 A1 E 2 A 2
F
+

1 =
A1
E1 A1 E 1 A1

E 1 A1 + E 2 A 2
F

1 =
E 1 A1
A1

F
E 1 A1

E 1 A1 + E 2 A 2
E 1 A1


1 =
E 1 A1
E 1 A1 + E 2 A 2

E 1 A1 + E 2 A 2 A1

F
E 1 A1 + E 2 A 2

1 =

Reduction Two: Input F, Output 1. This time we will start


with
F = F1 + F2
Eliminate F1 and F2, using the definition of stress

1 =

F1
A1

F1 = 1 A1

and 2 =

F2
F2 = 2 A2
A2

yielding

A2 1

A1

Was this the most efficient solution? Probably not, but


that is irrelevant. The point is that once you have written a
complete set of independent mathematical statements for all
relevant physical truths for a system, then you can solve for
any output variable.

E1

Now, check the equation. Always check units. If the units


of any term in the equation are wrong, then the equation is
wrong. These units check, since the unit of stress is force/
area. Now check for reasonableness, since the equation can
be wrong, even if the units are correct. This equation tells us
that the stress in bar one increases, if either the area of the
bar decreases, or the modulus of the bar increases. Is this
reasonable? Yes.

F = 1 A1 + 2 A2
Eliminate 2, using Hookes law 2 = E2 2 .
F = 1 A1 + E2 2 A2
Eliminate 2, using the definition of strain 2 =
F = 1 A1 + E2 A2

x2
.
L

x2
L

Eliminate x2, using x1 = x2 .


x1
L

F = 1 A1 + E2 A2

Eliminate x1, using the definition of strain 1 =

x1
.
L

F = 1 A1 + E2 A2 1
Eliminate 1, using Hookes law 1 = E11 .

1 = E11 1 =
F = 1 A1 + E2 A2

E1

E1

The only variables are the input F and the output 1. We are
finished substituting. Collect the output variable 1, and express the result in a useful form:

1.5 Overview of Engineering Modeling and Analyses

open
switch

35

+
C

Fig. 1.39 An electric circuit consisting of resistor R, inductor L,


capacitor C, a battery, and a switch

F = 1 A1 + E2 A 2

E1

other electrical circuit components are identified by their


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
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.
Step 1: Draw and annotate a picture Electrical schematics are lumped parameter models of electric circuits, so we

E2 A 2

A1 E1 E2 A 2
F = A1 +
+
1 F =
1

E1
E1

E1

A1 E1 + E2 A 2
A1 E1 + E2 A 2
E1
E1
F=
1
F=

1
E1
E1

A1 E1 + E2 A 2
A1 E1 + E2 A 2

1 =

E1
F
A1 E1 + E2 A 2

1.5.2.2Example 2: Electric Circuit, a Dynamic


System
An electric circuit consisting of a battery, a switch, a resistor with resistance R, an inductor with inductance L, and
a capacitor with capacitance C is shown in the schematic,
Fig. 1.39. Derive a differential equation which relates the
battery voltage to the voltage across the capacitor.
We will investigate the electrical phenomena and electric circuits in Chap.5. For the time being, we will work
with them as energetic systems presented as a network (i.e.,
in a circuit diagram). Current i is the quantity, which flows
through the network elements, and voltage v is the driving
potential, which drops in the direction of the flow. The terminology, driving potential, comes from electric circuits.
Electrical schematics are lumped parameter models of
electric circuits. The elements are identified by their parameter and by a graphic symbol. Components are modeled as
having a single electrical property. This is only a model. All
components possess the ability to dissipate and store energy.
Both electric and magnetic fields store energy. Electric fields
are created by electric charge. Magnetic fields are created
by electric charge in motion. Consequently, some energy is
stored in any an electric current. There is also some energy
lost with the flow of electric charge, which manifests as a
voltage drop in the direction of the current.
A component is modeled as a resistor, if its electrical
resistance is its dominant electrical property. Likewise, the

do not have to draw a picture. However, we do have to annotate the picture to show all of the variables and parameters
used in our equations. We must identify the nodes between
the elements. We must also show an assumed positive direction for the current flow through each of the elements. The
positive direction is arbitrary, except for the battery which
is the power source. The negative and positive terminals are
identified by the short and long bars of the pile symbol in
addition to the plus and minus signs. We will identify the
current through each individual element. Individual currents
are not necessary in this simple circuit, because all of the
elements have the same current flowing through them, since
there is only one loop. Currents flowing through an element
will be identified by the elements parameter as a subscript.
For example, the current through resistor R will be identified
as iR. The exception will be the current flowing through a
source which does not have a parameter.
There are two ways to denote the potential or voltage
drop between nodes in an electric circuit. One can either
use a subscript which indicates the element across which
the voltage drops, e.g., vR, to indicate the voltage dropping
across resistor R, or, alternatively, two subscripts to indicate
the nodes of distinct voltages between which the element
is connected. We will use the node notation that we introduced in Sect.1.4.2 using the piping system. Although the
node subscripts are more complicated than parameter subscripts, the node notation yields benefits when working with

1 Introduction to System Dynamics

36

closed
switch

+
i

iR

iL

iC

Fig. 1.40 The electric circuit of Fig.1.39 annotated with voltage nodes
between the elements, and the positive direction of current flowing
through the elements

complicated systems. Specifically, node subscripts eliminate variables and equations, when elements are connected
in parallel. Recall the convention introduced in Sect.1.4.2,
wherein the order of the subscripts is the order of the subtraction to calculate the difference in the driving potential. e.g.,
v1 v2 v12
Also recall that reversing the order of the subscripts inverts
the sign of the potential or voltage drop.
v12 = v21 since v1 v2 = ( v2 v1 )
The annotated schematic showing the nodes of distinct values of potential and the assumed positive direction for current through each element is Fig.1.40.
The two nodes identified as g for ground have the same
voltage because the conductor between them is modeled as
ideal. There is no voltage drop in the direction of current
flow through an ideal conductor. If two nodes always have
the same voltage, then they are, in fact, the same node, as
indicated by the dashed oval on the circuit diagram. Ground
voltage is an arbitrary reference voltage. It may or may not
correspond to earth ground, which is the voltage of the
earth at the location of the circuit. An alternative reference
voltage is chassis ground, where the chassis is, or used
to be, the mechanical support of the circuit components.
Chassis ground now commonly refers to the voltage reference of the power supply of the circuit. The ground reference
g in Fig.1.40 is the negative terminal of the battery; it is the
chassis ground.

compatibility equations we write, because we only want independent equations. The maximum number of useful compatibility equations equals the number of internal loops in
the circuit. Internal loops are the smallest loops; the holes in
the network. If we write more compatibility equations than
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.
There is one internal loop in this circuit. Using nodal subscripts, the compatibility equation is
v1g = v12 + v23 + v3g
2. Continuity Equations (Node Equations): These equations express the fact that, because nodes are geometric locations and not physical elements hence have no physical
properties, nodes cannot store electric charge. Current that
flows into a node must flow out. Once again, we use the sign
convention, and define flow into a node to be positive and
flow out of a node to be negative. We need one fewer node
equation than there are nodes in the system, to avoid creating a dependent equation. By convention, we will not write
a node equation for the ground node. The node equations for
this simple circuit do not convey much information, since
the current is the same through each element. In the general
case, however, the current though each element is unique.
Node 1: iSource = iR

Node 3: iL = iC

3. Elemental Equations: The elemental equation must respect the assumed positive direction of the current in the element indicated on the schematic. The potential or voltage
drop in the elemental equation is in the direction of the assumed current. First, write the expression for the type of element, and then edit it to describe the specific element in this
circuit, by adding the elements parameter as the subscript to
current, i, and the nodes the element is connected to as the
subscript to voltage, v. The node subscripts on voltage must
be ordered in the positive direction for the current through
the element, which is also the positive direction for a voltage
drop across the element.
Resistor: v = Ri

di
dt
dv
Step 2. Write mathematical statements of physical truth. Capacitor: i = C
dt
1. Compatibility Equations (Loop or Path Equations):
These equations express the fact that voltage drops around
loops must sum to zero. It makes no difference where you
start the summation. It does make a difference how many

Node 2: iR = iL

Inductor: v = L

v12 = RiR
diL
dt
dv3 g

v23 = L
iC = C

dt

4. Energy and Power Equations: There are two energy storage elements in this circuit, the capacitor and the inductor,
and one dissipative element, the resistor. We can write the
equation that the energy stored in the system is the sum of

1.5 Overview of Engineering Modeling and Analyses

the energy stored in the inductor and the energy stored in


the capacitor. We can also write equations for the amount of
energy stored in the inductor and in the capacitor. Finally, we
can write an equation for the power dissipated in the resistor.
In Chap.3, we will use energy equations to derive the initial
conditions we need to solve the differential system equation.
As we will explore in depth, the energy storage variables
define the energetic state of an energetic system, and will be
called the state variables.
Energy Stored in the System: E sys = E L + EC
1 2
LiL
2
1
Energy Stored in the Capacitor: EC = Cv32g
2
Energy Stored in the Inductor: E L =

Power Dissipated in the Resistor: PR = iR v12


5. Reduction and Solution. The only difference between the
reductions for the previous example of the elastic structure
and this circuit is that the elemental equations for the capacitor and inductor are differential equations. Consequently,
the resulting system equation will be a differential equation.
The order of the differential system equation will equal the
number of independent energy storage elements; two in this
case. We will return to the question of what makes an energy
storage element in a system independent. For the moment,
when the two energy storage variables are different types of
variables, current iL and voltage v3g, as in this case, then the
energy storage elements are independent.
There are three firm rules to follow during the reduction
of a set of equations a differential system equation:
1. Do not use the energy equations to derive the system
equation. Energy equations are integrals. We are deriving
a differential equation. The energy equations are used to
provide the initial conditions needed to solve the differential system equation, but not to derive the equation.
2. Never integrate during the reduction, only differentiate.
The system equation is a differential equation, not an integral equation.
3. If you reach the end of the reduction, and you have not
introduced all of the elemental parameters, then there is
an error.
The reduction technique is elimination by substitution. The
only difference we will see from the previous mechanical
example is that we may have to differentiate both sides of the
equation with respect to time, in order to make a substitution.
We can randomly pick any equation from our set of
equations, except a trivial equation, and start substituting to
eliminate all unwanted variables. Our input variable is the
voltage across the battery, v1g, and the output variable of interest is the voltage across the capacitor, v3g. It is generally
easiest to start with an equation that contains either the input

37

or the output variable. We will start with the compatibility


equation:
v1g = v12 + v23 + v3g
We keep the input and output variables, v1g and v3g. We must
eliminate v12 and v23. We can substitute the resistor elemental
equation for v12.
v1g = RiR + v23 + v3 g
Now we must eliminate iR. We use the continuity equation
for node 2, iR = iL , to eliminate iR:
v1g = RiL + v23 + v3 g
and the continuity equation for node 3, iL = iC , to eliminate
iL:
v1g = RiC + v23 + v3 g
Introduce the derivative of the output variable v3g by substituting in the capacitors elemental equation
iC = C

dv3 g
dt

to eliminate iC.
v1g = RC

dv3 g
dt

+ v23 + v3 g

Voltage v23 is eliminated, using the inductors elemental


equation:
v23 = L

diL
dt

yielding
v1g = RC

dv3 g
dt

+L

diL
+ v3 g
dt

Current iL is eliminated, by again using the continuity equation for node 3, iL = iC ,


v1g = RC

dv3 g
dt

+L

diC
+ v3 g
dt

Current iC is eliminated, by again using the element equation


for the capacitor, iC = C
v1g = RC

dv3 g

dv3 g
dt

dt

+L

d dv3 g
C
+ v3 g
dt dt

1 Introduction to System Dynamics

38

We have finished substituting. We now apply the differential


operator

dv3 g
d
, yielding
to C
dt dt
v1g = RC

dv3 g
dt

+ LC

d 2 v3 g
dt 2

+ v3 g

This is a second-order differential system equation. Put it


into standard form for higher-order differential equations, by
arranging the terms in order of decreasing differentiation and
clearing the coefficient of the highest-order derivative.
v1g = LC

d 2 v3 g
dt

+ RC

dv3 g
dt

Distribute the units of operator onto the terms of the summation on the right side.
2

d v3 g R dv3 g 1
1
=
v
LC 1g dt 2 + L dt + LC v3 g

All of these terms must have the same units as the single
term on the left-hand side. Second-grade math applies to
differential equations. You cannot add apples and oranges.
First, simplify the expressions, by dropping the subscripts on
the variables, and by evaluating the units of the derivatives.
1 v R v 1
LC v = t 2 + L t + LC v

+ v3 g

d 2 v3 g R C dv3 g
1
1
+
v1g =
v3 g
+
2
LC
LC
dt
L C dt
d 2 v3 g R dv3 g
1
1
v1g =
+
+
v3 g
LC
L dt
LC
dt 2
We will review the method of undetermined coefficients
and the use of Laplace transformations for solving ordinary
differential equations with constant coefficients in Chap.2.
However, before we invest time in solving an equation, we
will check the consistency of its units. If the units are inconsistent, then the equation is erroneous. Remember that
checking for unit consistency only reveals some errors. The
equation can still be wrong, even if the units of each term are
consistent.
d
What are the units of the derivative operator, ? Recall
dt
when the derivative was introduced in Calculus I. The operator d represents an infinitesimally small change. It has no
1
d
are . Square brackets
units. Consequently, the units of
t
dt
are used to denote the units of, so the previous sentence
can be expressed symbolically as
d 1
(1.48)
dt = t
Similarly, the units of the second derivative with respect to
time are

What remains is an expression consisting of the variables,


v and t,and the elemental parameters. If our purpose was a
dimensional analysis, then we would express the parameters,
R, L, and C, in terms of fundamental units. However, it is
not our purpose to perform a dimensional analysis. We simply wish to check the consistency of the units of each term
in the equation. We will find that performing a dimensional
analysis becomes a challenge, when we work with motors
and other hybrid systems which contain two or more different types of energetic subsystems.
The easy way to check the consistency of the units of the
system equation is to express the units of the elements parameters in terms of the power variables of the system and
time. By rearranging the element equation for the resistor,
inductor, and capacitor to solve for the parameter, and then
applying the units of operator [ ]:
Resistor:
v
v
(1.50)
v = Ri

R=

[ R ] = i
i

Inductor:
di
dt
vt
(1.51)
v=L
L=v
[ L] =
dt
di
i
Capacitor:

d 1
(1.49)
2= 2
dt t

dv
dt
it
(1.52)
i=C
C=i

[C ] = v
dt
dv

The units of the differential system equation are checked by


applying the units of operator [ ] to both sides of the equation.

We have expressed the units of the element parameters in


terms of the electrical systems power variables, i and v, and
the independent variable time t. Substituting the units of the
parameters into the differential equation:

1
d v3 g R dv3 g
1
=
+
+
v
v3 g

1
g
2
LC
L dt
LC
dt

1 v R v 1
LC v = t 2 + L t + LC v

Summary

39

yields
v
i v v i
v = 2 +

vt it t vt
i

v i v
v
+
t vt it

Cancel terms, but be careful with improper ratios.

i v v
v = 2 +

v t it t

v i

i v v t + i v v

v t t i v t it

i
v t

i v v v v i i v
v t it v = 2 + i t v t + v t it v

i v v v v i i v
v = 2 +
v

+
v t it t i t v t v t it
v v v v
t 2 = t 2 + t 2 + t 2
The units of each term are consistent.

Summary
Chapter 1 introduced system dynamics as the study of energetic systems; systems in which the flow of energy from
a power source into and out of energy storage and dissipation elements produces a dynamic response of the power
variables of that system. Energy is a quantity which can
move, be stored, change form, and be dissipated as heat.
Mechanical energy storage elements, the ideal spring which
stores elastic strain energy, and the ideal mass were used to
introduce energy and coenergy. Ideal springs and masses
were also used to introduce the concept of energetic models
in which an energetic element possesses a single property
or attribute. Modeling of physically real systems which are
made of components that possess a number of different energy properties is the subject of the remainder of this text.
The process of engineering modeling and analysis was
also discussed. It is an iterative process. It begins with a
simple model of the part, machine, or system. The models
complexity is increased, as insight is gained from the results
of the previous iteration, until one of the following occurs:
You have the information you need to make the engineering decision, or
You have run out of time or money, or

The accuracy or reliability of the model or the data the


model uses has reached its useful limit.
Engineers are typically more visual than verbal people. Most
engineering design is fundamentally graphical. Although
practicing engineers need to understand calculus and differential equations to understand the physical processes and
design methods we use, it is more important to be able to
visualize the mathematical relationships than it is to remember formulas. The vast majority of the mathematics done by
practicing engineers is arithmetic with a small amount of
algebra. The primary challenge of engineering mathematics
is organizational not conceptual. If you have been relying
upon example problems in texts rather than understanding
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
represented graphically. Engineers need pictures. Learn how
to use them to your advantage.
The steps in engineering analysis are to:
1. Draw a reasonable model, annotated to show the relevant
variables and parameters. Never write an equation which
contains a variable or parameter that has not been defined
on a graphic.
2. Write a complete set of independent mathematical statements of physical truth relevant to that model.
3. Reduce the mathematical statements to relationships
which relate what we know (the input) to what we want to
know (the output).
4. Evaluate the solution. Iterate if necessary.
Begin your transition from engineering student to practicing
engineer now.
A model represents only one (or some) of the attributes of
the subject object, system, or process.
The model is the basis of the subsequent calculations.
Therefore, it is of utmost importance to be confident that
the model represents the actual physical object, system,
or process with sufficient accuracy for the intended use of
the analytical results.
Engineering models can be physical, analogous, or mathematical. It is important to include a graphical representation of the model, and to update it as the model evolves.
Calculations yield results. There are no answers in the
sophomoric sense, because the results vary as the model
is refined.
Every engineering calculation requires interpretation, i.e.,
What does this result mean?
Although analysis is challenging for an engineering student to learn, synthesis is more challenging for the practicing engineer because the choices are open ended. Analysis is algorithmic. Design is synthesis.
Increasing understanding, knowledge, and experience
increases the design options of the engineer.

1 Introduction to System Dynamics

40
Fig. P1.1a Force F(t) acting on
a mechanical system. b Colinear
velocity v(t)

200
150

F(t)
N

v(t)
m
___
sec

100
50
0

2.0

-50

t, sec

1.0
0.5
0

1.5

t, sec

-100

Fig. P1.2a DC current i(t) driving a motor. b Voltage v(t) of the


power supply

a
i(t)
A

20

200

15

150

10

v(t) 100
VDC

5
0

50
0

-5

t, sec

-50

-10

-100

Problems
Problem 1.1Translational 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
point of application of the force are colinear, the scalar equation P = F v can be used. A mechanical system was de-energized before the force shown in Fig.P1.1a with the colinear
velocity shown in Fig.P1.1b acted on it.
1.1.a Plot the energy provided by the power source from 0
to 4sec.
1.1.b What is the minimum horsepower motor which could
provide the power n eeded?
Problem 1.2DC Electrical power is product of current
times voltage, P = iv. A DC motor and the system it drives
was de-energized, before the DC motors power source provided it the current i at the voltage v, shown in Fig.P.1.2a
and b.
1.2.a P
 lot the energy provided by the power source from 0
to 8sec.
1.2.b What is the minimum horsepower motor which could
accept the power from source?
Problem 1.3A piping system is shown in Fig.P1.3. The
fluid is modeled as incompressible. Consequently, the continuity equations can be written in terms of volume flow rate,
Q, rather than in terms of mass flow rate. The branches are
identified by letter, and pressure nodes between the branches
are numbered.

t, sec

Branch A 2

Branch B

1
Pump

Branch C
patmosphere

Branch D

Branch E
Fig. P1.3 Piping system schematic. The fluid is modeled as incompressible. There is fluid resistance in each branch, which decreases
pressure in the direction of the fluid flow

1.3.a O
 rient the flow in each branch. The positive direction
for flow through pump is from node 5 to node 1.
1.3.b Write a complete set of independent compatibility
equations in the form of path equations.
1.3.c Write a complete set of independent continuity equations for volume flow rate.
Problem 1.4A piping system is shown in Fig.P1.4. The
fluid is modeled as incompressible. Consequently, the continuity equations can be written in terms of volume flow rate,
Q, rather than in terms of mass flow rate. The branches are

Problems

41

Branch A 2

Branch B

80,000

patmosphere

60,000

1
Pump

Branch C

Branch D

F, lb
40,000
20,000

Branch E

Branch F
Fig. P1.4 Piping system schematic. The fluid is modeled as incompressible. There is fluid resistance in each branch, which decreases
pressure in the direction of the fluid flow

x, in

Fig. P1.6 Force in pounds vs. displacement in inches. The test results
x

are approximated by F (t ) = 7 104 lb 1 e in

yield

closed
switch

+
i source

Stress,

iR

iL

iC

g
yield

Strain,

rupture

Fig. P1.5 Stressstrain plot of a specimen tested to rupture. The specimen was 0.5in. in diameter and 3in. long. The yield stress, yield strain,
and rupture strain are yield = 90ksi yield = 0.18% rupture = 1.02%

identified by letter, and pressure nodes between the branches


are numbered.
1.4.a Orient the flow in each branch. The positive direction
for flow through pump is from node 5 to node 1.
1.4.b Write a complete set of independent compatibility
equations in the form of path equations.
1.4.c Write a complete set of independent continuity equations in terms of volume flow rate, Q.
Problem 1.5 The stress vs. strain plot of a specimen tested
to rupture is approximated by the elasticperfectly plastic
model shown in Fig.P1.5.
1.5.a Calculate the energy density needed to break the specimen in US customary and SI units.
1.5.b How much energy was needed to break the specimen?

Fig. P1.7 Electric circuit consisting of a battery, resistor, inductor,


and capacitor in series, annotated with voltage nodes and the positive
directions of current through the elements

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
the displacement, x = 4 in., in foot-pounds and joules.
1.6.b Calculate the coenergy for the displacement, x = 4 in.
Problem 1.7An electric circuit consisting of a voltage
source, represented as a battery, a switch, a resistor, an inductor, and a capacitor, is shown in Fig.P1.7. The energetic
equations of this circuit are listed. Use elimination by substitution to derive the system equation for the input voltage and
output variable indicated:
i. Input Variable: v1g, Output Variable: v12
ii. Input Variable: v1g, Output Variable: v23
iii. Input Variable: v1g, Output Variable: v3g
iv. Input Variable: v1g, Output Variable: iL
and check its units in terms of the power variables, current,
voltage and time.

1 Introduction to System Dynamics

42

i source

iL

iR

L iC

iL

i source

R iC

iR

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

Fig. P1.9 Electric circuit consisting of a battery, resistor, inductor, and


capacitor in series, annotated with voltage nodes and the positive directions 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 Eq:
v1g = v12 + v23 + v3g
Element Eqs:

Capacitor: iC = C

v1g = v12 + v2 g

v2 g = v2 g

Element Eqs:
Resistor: v12 = RiR

Resistor: v12 = RiR


Inductor: v23 = L

Compatibility of Voltage Drops, Path Eqs:

diL
dt
dv3 g
dt

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

Inductor: v2 g = L
Capacitor: iC = C

diL
dt
dv2 g
dt

Energy Eqs:
System: E sys = E L + EC
1 2
LiL
2
1
Capacitor: EC = Cv22g
2
Inductor: E L =

Problem 1.9 An electric circuit consisting of a voltage


source, a resistor, an inductor, and a capacitor is shown in
the schematic Fig.P1.9. The energetic equations of this circuit are listed. Use elimination by substitution to derive the
system equation for the input and output variable indicated:
i. Input Variable: v1g, Output Variable: v12
ii. Input Variable: v1g, Output Variable: v2g
iii. Input Variable: v1g, Output Variable: iL
iv. Input Variable: v1g, Output Variable: iR
v. Input Variable: v1g, Output Variable: iC
and check its units in terms of the power variables, current,
voltage and time.

Problems

43

Fig. P1.10a Force F(t) acts


on mass M which slides on a
lubricating film with damping b
against spring K which is connected to ground. b The linear
graph of the system

x,v

F(t)
g

F(t)

1
g

Lubricating fluid
Damping b

Energetic Equations:

Continuity (Conservation of Charge), Node Eqs:


Node 1: isource = iL

Pump

Element Eqs:
Resistor: v2 g = RiR
diL
dt
dv2 g
Capacitor: iC = C
dt
Inductor: v12 = L

R2
patm

Fluid
Reservoir atm

Compatibility of Voltage Drops, Path Eqs:


v2 g = v2 g

R1

p(t)

vent to
atmosphere

Node 2 : iL = iR + iC

v1g = v12 + v2 g

Fig. P1.11a Fluid system modeled as a pressure source, two fluid resistances, a fluid inertance, and a fluid capacitance

Energetic Equations:
Continuity (Force Equilibrium) Node Eqs:
F (t ) = Fb + FM + FK

Energy Eqs:

Compatibility of Velocity, Path Eq: v1g = v1g

System: E sys = E L + EC

Element Eqs: Fb = bv1g

1
Inductor: E L = LiL2
2
1
Capacitor: EC = Cv22g
2
Problem 1.10 A translational mechanical system consisting
of a mass M sliding on a lubricating fluid film with damping
b, and a spring K attached between the mass and ground is
shown in Fig.P1.10a. The linear graph of this energetic system, analogous to an electric circuit diagram is Fig.P1.10b.
The energetic equations are listed. Use elimination by substitution to derive the system equation for the input and output
variable indicated:
i. Input Variable: F(t), Output Variable: Fb
ii. Input Variable: F(t), Output Variable: FM
iii. Input Variable: F(t), Output Variable: FK
iv. Input Variable: F(t), Output Variable: v1g
and check the its units in terms of the power variables, force,
velocity and time.

FM = M

dv1g
dt

dFK
= Kv1g
dt

Energy Eqs:
System: E sys = E M + E K
1
Mv12g
2
F2
Spring: E K = K
2K
Mass: E M =

Problem 1.11 A schematic of a hydraulic system is shown


in Fig.P1.11a. The pump, modeled as a pressure source p(t),
discharges fluid into a hydraulic circuit consisting of two
fluid resistances, R1 and R2, a fluid inertance I, which stores
kinetic energy, and a fluid accumulator with capacitance C,
which stores energy by compressing a spring or nitrogenfilled bladder. FigureP1.11b is the linear graph of the system, analogous to an electric circuit. The energetic equations

1 Introduction to System Dynamics

44

R1

Element Eqs:

Fluid Resistance R1: p12 = R1QR1


Fluid Resistance R2 : p3 g = R2 QR2

R2

p(t)

patm

patm
Fig. P1.11b Linear graph of the hydraulic system

are listed. Use elimination by substitution to derive the system equation for the input and output variable indicated:
i. Input Variable: p(t), Output Variable: p12
ii. Input Variable: p(t), Output Variable: p23
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
and check its units in terms of the power variables, pressure,
volume flow rate Q, and time.
Energetic Equations:
Continuity (Conservation of Volume of Incompressible
Fluid), Node Eqs:
Node 1: Q = QR1
Node 2: QR1 = QI
Node 3: QI = QR2 + QC
Compatibility of Pressure Drops, Path Equations:
p1g = p12 + p23 + p3g

p3 g = p3 g

Fluid Inertance I : p23 = I

dQI
dt

Fluid Capacitance C: QC = C

dp3g
dt

Energy Eqs:
System : E sys = E I + EC
1 2
IQ I
2
1
Capacitance: EC = C p32g
2
Inertance: E I =

References and Suggested Reading


Crandall SH etal (1982) Dynamics of mechanical and electromechanical systems. Krieger, Malabar
Karnopp DC, Margolis DL, Rosenberg RC (2012) System dynamics:
modeling, simulation, and control of mechatronic systems, 5th edn.
Wiley, New York
Kulakowski BT, Gardner J, Shearer L (1997) Dynamic modeling and
control of engineering systems, 3rd edn. Cambridge, Cambridge
Ogata K (2003) System dynamics, 4th edn. Prentice Hall, Englewood
Cliffs
Rowell D, Wormley DN (1997) System dynamics: An introduction.
Prentice Hall, Upper Saddle River
Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
dynamics. Addison-Wesley, Reading

Differential Equations, Input Functions,


Complex Exponentials, and Transfer
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 superposes or sums the response of a system into the natural or homogeneous response of the system 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 Eulers 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 chapters appendix introduces Mathcad
and MATLAB to plot the solutions or response functions.

2.1Introduction
This chapter presents most of the advanced mathematics used
in the remainder of the text. The topics and methods introduced here will be developed further, when they are applied
in later chapters. Most engineering students do not understand advanced mathematics, until applying it to physically
meaningful problems. This is particularly true for differential equations. It is certainly true that mathematical ability
varies between individuals, and that understanding advanced
mathematics takes some students longer than others. However, with persistence, it is possible for all engineering students
who have passed their courses in calculus and differential
equations to gain an understanding of this mathematics.
Any quantity which flows requires time to move or accumulate. Energy flows across system boundaries and between elements within an energetic or dynamic system. The
time dependency of energy storage leads to the power flows
in dynamic systems described by differential equations. The
interaction of the individual energetic elements which comprise a system with the power source that provides the input is
described by differential system equations. The solution of a

differential system equation for an input is a prediction of the


response of the actual system to that input, Fig.2.1.
The most accurate description of dynamic systems requires non-linear differential equations. However, most nonlinear differential equations cannot be solved manually using
classical techniques. We can solve non-linear differential
system equations using the appropriate numerical method,
such as RungeKutta method, if the solution exists, and we
will learn how to use the computational tools, Mathcad and
MATLAB, to solve linear and non-linear differential equations.
However, engineers must also understand the analytical
solutions of ordinary differential equations, in order to understand the physics of dynamic systems, and to effectively use
design techniques developed from the analytical solutions.
We shall apply an engineering perspective to differential equations. The differential equations we will derive
are inputoutput relationships. These differential system
equations will be viewed as operating on the input to the
energetic system, an input that is under our control, to yield
an output variable, any power variable in the system, whose
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

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

46

Dynamic
System

Input Variable
Forcing Function

Output Variable
General Solution

Fig. 2.1 Differential system equations are inputoutput relationships


used to predict the response of energetic systems to different inputs

Unit Step Input

F(t)

v(t)

Unit Step Response

1
0

time

Input Variable, F(t)

Differential
System
Equation

time

even when that knowledge is far from complete, to make


rational decisions and recommendations.
We will limit ourselves to linear differential equations
and view them as dynamic operators which operate on an
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
are time histories of the systems power variables. An understanding of the mathematics in this chapter changes a
students perspective from How do I solve this differential
equation? to What is the possible behavior in this system?

Output Variable, v(t)

Fig. 2.2 A differential system equation can be viewed as a mathematical operator. The differential system equation operates on an input function to yield the corresponding response function

We shall see that, if we restrict the models of the energy


storage and dissipation mechanisms, such that they (1) have
constant properties, and (2) a linear relationship between
two power variables, or between a power variable and the
time derivative of a power variable, the resulting differential
system equation is linear. Linear differential system equations allow the use of superposition, in which a solution is
constructed by breaking down an arbitrary input to a sum of
simpler inputs, solving for each input individually, and then
summing the solutions. This is the basis of the Method of
Undetermined Coefficients.
We will extend our use of superposition to construct arbitrary inputs to our dynamic systems from scaled, timeshifted, and superimposed Heaviside unit step functions.
The response of a dynamic system to an arbitrary input is
constructed by performing the same scaling, time shifting,
and superimposing of the response of the system to a single
unit step input, as was performed on the input unit step functions. The use of unit step response functions will shift
our focus from how to solve linear differential equations to
how to use the solutions of these equations in engineering
applications, Fig.2.2.
The carrot for the engineering student, who believes that
he or she never understood differential equations and cannot bear to look at another differential equation, is straightforward. Step response functions allow us to identify the
solution of differential system equation, rather than solve
the equation. There is a less obvious benefit also. Textbook
problems provide sufficient information to apply rigorous
mathematical methods. The nature of the design process is
the opposite. Although practicing engineers must anticipate
and accommodate the response of an energetic system to a
power input, they rarely have the luxury of formulating a
differential equation. Typically, very little is known, since
the design is still in flux. An engineer must use the available
analytical tools with what he or she knows about a system,

2.2 Input Functions


The inputs to differential system equations are mathematical functions which approximate the time history of a power
variable of an actual power source driving a physical system. The most common input functions are singularity
functions and sinusoids. A singularity is a finite quantity divided by zero. The useful singularity input function
is the Heaviside unit step function. The singularity of the
Heaviside step function is its time derivative at the instant
the function makes it step change from zero to one. The
step change is modeled as occurring instantaneously. Hence,
there is a finite change in zero time. The two other singularity functions commonly used as input functions are the unit
impulse function and the unit ramp function, which are the
time derivative and the time interval of the Heaviside step
function, respectively. Sinusoids are used to represent periodic inputs. The Fourier series allows any periodic function
to be approximated by a sum of scaled (or weighted) sine
and cosine terms of integer multiples of a base or fundamental frequency. The use of a single sinusoid as an input is
presented in this chapter. The Fourier series and frequency
response, which is the steady-state response of a system to
a sinusoidal input, are discussed in Chap.10.

2.2.1 Power Sources


An input to an energetic system is one of the two power variables of a power source connected to the system. We model
an ideal power source as a device which can maintain a value
of one of its two conjugate power variables by supplying as
much power as is drawn by the system. We can only control
one of the two power variables of a power source. The power
drawn from the source by the system determines the value of
the second power variable.
There are no ideal power sources, but there are many that
can be reasonably approximated as ideal. In general, the less
power drawn from a power source relative to its capacity,
the more ideally the power source behaves. For example, a

2.2 Input Functions

47

chemical battery, such as an automotive lead-acid storage


battery, will maintain a fixed voltage over a range of currents
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 batterys capacity. When the current drawn is no longer relatively small, we generally need to
include an energy dissipater in the source model, so that the
output voltage decreases as the current flow increases. The
usefulness of any dynamic model is limited to an operating
range of the systems variables.
An ideal linear power source must source, or provide
power, and sink, or accept power. A lead-acid storage battery provides power when it discharges and accepts power
when it recharges. Many batteries can provide only power.
They cannot be recharged. We deal with a source that cannot
accept power, by limiting the valid range of the results of the
dynamic model. The results are valid, until the time the model
predicts the power flowing back into the source. If that duration
of the response is too brief to reveal the aspect of the system
we are interested in, then we must program a computational
model in order to impose conditional logic on the behavior of
the source and solve it using a numerical method. Numerical
methods are computer-based solutions using finite differences
to approximate derivatives. They are presented in Chap.8.

2.2.2 Heaviside Unit Step Function


A typical differential equations problem reads as follows:
6

dv
+ 3v = 8
dt

which, if interpreted literally, makes no physical sense, because the input is not a function of time. It is common mathematical notation, however, because it is implied that the input
is applied to the system at time, t = 0 , where time, t = 0, is
defined as the beginning of the dynamic response. For our
analyses to be useful, we need to be able to turn on and turn
off the input when we wish. The Heaviside unit step function,
invented by the British engineer, Oliver Heaviside, serves
that purpose, Fig.2.3.
The Heaviside unit step function has two possible values.
It equals zero, when its argument is negative, and equals one,
when its argument is positive. It is defined as:


us (t ) = 0 for t < 0
us (t ) = 1 for t 0

(2.1)

where u stands for unit magnitude and the subscript, s


stands for step. We will use this notation. The Heaviside step
functions change from zero to one is referred to as its transition or, informally, when the step turns on. It is important

us(t)
1

time

Fig. 2.3 The Heaviside unit step function. The step function transitions, or changes its output value fromzero to one, when its argument
equals zero

to note that the Heaviside step function is constant, except at


the instant it transitions from zero to one. It has just one transition. Consequently, the Heaviside unit step function cannot
be turned off. It must be negated by adding or superposing a negative step at the time we wish the sum to equal zero.
We will address time shifts in Chap.3.
There is neither a single definition nor standard notation
for the Heaviside unit step function. The Heaviside unit step
function is sometimes defined to have a value of 0.5 when
the argument equals zero:


us (t ) = 0

for

t<0

us (t ) = 0.5

for

t=0

us (t ) = 1

for

t0

(2.2)

The most common notation for Heaviside is H (t ). Another common notation is 1(t ), where 1 is the number one.
Mathcad uses (t ) . MATLABs symbolic math toolbox uses
Heaviside(t).
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
instant the step input is applied. The initial conditions are
usually presented in textbook differential equation problems at time, t = 0 . Since the transition of the Heaviside
step function from the value of zero to the value of one
is instantaneous, we need to split hairs and define three
different time zeros. Time, t = 0, is the instant the step
is applied. We will call the instant before the step is applied, t = 0 , and the instant after the step is applied, t = 0+ ,
Fig.2.4.
The derivative of the Heavisides step function transition
is a problem. Mathematically, the value of the derivative at
the instant of the transition is infinite, because the transition
from zero to one occurs instantaneously, that is, over zero
time. The derivative of the step function is zero for all other
values of time, since the step function is constant, except at
the instant of its transition.


dus (t )
= for t = 0
dt

(2.3)

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

48

us(t)
1

t>0
us(t) = 1

(t)

t<0
us(t) = 0

t<0
(t) = 0

t>0
(t) = 0

time
t = 0-

t=0

time

t = 0+

t = 0-

Fig. 2.4 The Heaviside unit step function, us (t ) , Eq.2.1. The time,
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

and


dus (t )
= 0 for t 0
dt

t=0

t = 0+

Fig. 2.5 The unit impulse function (t ), Eq.2.5

tial conditions needed to solve differential equations, after


we introduce state variables, and investigate the mathematical implications they have as the energy storage variables of a system.

(2.4)

If a Heaviside step function is differentiated as an input term


of a system, which value should we use, zero or infinity?
When the Heaviside step input is applied at time, t = 0, we
restrict the solution of the equation to time, t > 0, since the
input must precede the response to respect cause and effect, and use the value dus (t ) / dt = 0 . If the system is deenergized before the step input is applied, and the input term
of the system equation is differentiated, what drives the response of the system? Although the derivative of the step
input is zero, the input to the system is not zero. The input
to the system is a step. It is the input to the system, which
determines the initial values of all the power variables in the
system and their derivatives. The name, system equation,
is somewhat misleading. A more descriptive but awkward
name would be the differential equation for a power variable in an element in the system. There are two power variables in every element of a dynamic system. We can formulate a differential equation to describe the response of all but
one of these variables to the input applied to the system, with
the input variable itself being the only exception. The input
term (or terms) applies to the differential system equation of
a power variable of one energetic element of the system. The
differential system equation for the other power variable of
the same energetic element will have the same input, but the
input term(s) will be different; they will have different coefficients and differentiation.
We consider the Heaviside step function further in
Chap.3, when we discuss modeling input steps, and using
superposition (summation) of step inputs to form pulse inputs. We will also establish how to determine the initial
values and derivatives of the power variables, i.e., the ini-

2.2.3 Unit Impulse


The unit impulse function, Eq.2.5, Fig.2.5, is the derivative
of the Heaviside unit step function. The amplitude of the unit
impulse function is infinite, but its duration is infinitesimal


(t ) = for t = 0
(t ) = 0 for t 0

(2.5)

The unit impulse function is the derivative of the Heaviside


step function. Inversely, the integral of the unit impulse is
the unit step function. Hence, the area under the unit impulse equals one, which led to the name unit impulse,
when the impulse has infinite magnitude. A second common name for the unit impulse is the Dirac delta function,
(t). Paul Dirac was a British physicist influential in the
development of quantum mechanics. We will use the combination of the name unit impulse and the symbol, (t),
as a reminder that the magnitude is not unity. Needless to
say, it is impossible to apply a unit impulse with infinite
amplitude, but no duration, to a physical system. The unit
impulses use is purely mathematical. It represents the derivative of the Heaviside step function. The energy which
drives the response is delivered to the system by the step
function, not by impulse.

2.2.4 Unit Ramp


The integral of the Heaviside unit step is the unit ramp function, which has a slope or ramp rate of one. The unit ramp

2.3Linearity

49
1.0

ur(t)
1

sin(t) us(t)
0.5

time

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,


and the Heaviside step function, Eq.2.5, Fig.2.6.


ur (t ) = tus (t )

(2.6)

2.2.5Sinusoids
The other important type of input is a sinusoid. Sinusoidal
inputs occur naturally and are created by rotating machinery
and vibrations or oscillations within the equipment or physical system, which forms the environment around the system
of interest. Sinusoids are also the basis of the Fourier transformation, which allows any periodic input to be approximated
by a summation of sinusoids to the accuracy desired. A very
important special case is the steady-state response of systems
to a sinusoidal input, which is applicable to both vibration
analysis and electrical systems powered by alternating current. Mechanical vibrations are introduced in Chap.4. We
shall investigate the Fourier transform and the steady-state
response of systems to sinusoidal inputs in Chap.10. In this
chapter, we will consider the transient period up to steadystate.
In order to turn on a sinusoid at time, t=0, it is multiplied by the Heaviside unit step function per Eq.2.7:


f (t ) = sin (w t )us (t )

(2.7)

where is the angular frequency in radians per second.


Multiplication by the Heaviside step function is not necessary, if the problem is a textbook problem. It is necessary,
even for a textbook problem, when the sinusoid is applied at
a time other than t=0, Fig. 2.7.

2.2.6 Step Responses as Input Functions


Modeling a power source as an ideal source described by a
Heaviside unit step function is a mathematical convenience
which is always in error to some degree. All power sources
are physical devices with multiple energetic attributes. How

Fig. 2.7 The product of a sine and the Heaviside unit step function
f (t ) = sin (w t )us (t ), Eq.2.7

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
the characteristic times of both the power source and the
system. When the functional form of the step input from an
actual source deviates from the ideal, square-cornered step
described by a Heaviside unit step function, to the extent
the input cannot be reasonably modeled by the ideal step,
then the engineer modeling the system has two choices. The
system boundary can be expanded to include the power supply as a dynamic element within the model, and the system
model revised to include the previously unmodeled dynamic
aspects of the power source. Alternatively, if there is not
enough information to create a model of the power source
or to revise the system model, then the model of the input
must be changed. If the time history of the input variable has
been measured, then a step response function can be used
to model the input of the system. We will defer use of step
response functions as inputs until the introduction of transfer functions in Sect.2.10.

2.3Linearity
Many engineering calculations rely on superposition. The
premise of superposition is that the response of a system to
each piece of a combined input can be calculated separately,
and, then, the responses summed to calculate the response to
the entire input. In other words, superposition is the decomposition of an input into component parts, calculation of the
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
an input by summing different input functions. The hard part
is ensuring that the sum of the responses calculated for each
component of the input is reasonably close to the response of
the system to the entire input. Superposition (or summation)
of the resulting step responses only has validity, if we have
restricted our model to linear elements.

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

50
Fig. 2.8a Elasticperfectly
plastic material model. b Forces,
F1 and F2, superposed on a machine element loaded in tension

y
E

F=F1 +F2

F=F1 +F2
Area A
E, y

iR

30

v12
VDC

iR

20

100 x R

10

Fig. 2.9 The schematic symbol of electrical resistance showing nodes


of distinct values of voltage at either end of the element

-0.4 -0.3 -0.2 -0.1

0.1 0.2

-10

Then we can use superposition to calculate the response


of the system to an input comprised of step functions.
Although superposition is a fundamental tool in engineering analysis, it is often misused. Superposition works only if
phenomenon is reasonably approximated by a linear model
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 calculation of the stress in a rod due to two axial forces, F1, applied
at time, t = ta , and F2, applied at time, t = tb, Fig.2.8b. We
can use superposition to calculate the axial stress in the rod
by calculating the stress due to each force separately and
summing the stresses, thusly,

1 =

F1
A

and

2 =

F2
A

if and only if

1 + 2 = axial

axial y

since the maximum stress possible in the material model is


y. In other words, we must stay within the linear region of
the elasticperfectly plastic material model, if we wish to use
superposition. The result axial > y is nonsense when the
elastic-perfectly plastic material model is used, although it
is commonly observed in practice since many real materials
strain harden.
We will use the terms element model or constitutive
model for the equations which describe the relationship between two variables in an element of an energetic system.
For example, the familiar equation of electrical resistance,
v12 = RiR , is a linear element equation. Voltage v12 is the difference or drop in the voltage across the resistance in the
direction of the current, iR, flowing through the resistance.
The resistance, R, is a property of the element, Fig.2.9.
The linear element model for an electrical resistor is a
significantly simpler model than the elasticperfectly plastic
material model. The linear electrical resistance model is a

0.3 0.4

i, amps

-20
-30

Fig. 2.10 Linear electrical resistance. The slope of the line is the
resistance, R, when the units are volts and amperes

single straight line which passes through the origin of the


plot of the power variables, v12 and iR. Conversely, the elasticperfectly plastic model is comprised of two straight lines.
Consequently, it cannot be a linear model. It is a non-linear
model. Linear models must have a single straight line which
passes through the origin, Fig.2.10.
In addition to restricting our models of energetic systems
to linear models, we must also use linear mathematical operators, in order to use superposition. We can establish whether
or not a mathematical operation is linear with two simple
tests:
1. Can the operation by distributed onto the individual terms
in a sum?
2. Does doubling the input to the operation double the result?
Since we intend to derive a differential system equation, we
should establish the conditions needed for differentiation
with respect to time as a linear operation. Can the operation,
d , be distributed onto the terms of a sum, such as x + x ?
a
b
dt


d ( xa + xb ) dxa dxb
=
+
dt
dt
dt

(2.8)

Yes. Does doubling the input to the differential operator


double the result?


d (2 x)
dx
=2
dt
dt

(2.9)

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 example, define x xa2. Differentiating 2 x 2 xa2 with respect
to time yields the following:

( )

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

We will ensure that our dynamic models of energetic systems are linear, by using only linear equations to describe
the individual elements within the system, so that we can use
superposition. We must always remember to interpret our
results to evaluate the magnitude of the inaccuracy we introduce by the linear approximation of a non-linear physical
relationship. We will develop models of mechanical system
elements in Chap.4. For the time being, we will work with
models as presented.

2.4Superposition
The general solution of linear ordinary differential equations
with constant coefficients is the superposition or sum of the
solution to the homogeneous equation, created by setting the
output variable side of the differential equation to zero, and of
the particular solution(s) for each forcing function. The forcing functions are the inputs to the system. When an energetic
system has reached steady-state, there is equilibrium between
the energy storages in the system, Fig.2.11. The solution of
the homogeneous equation represents the natural response of
a system to a disturbance. If the system is then disturbed by
an input of energy, a new equilibrium must be reached.
The homogenous response of a dynamic system describes
how the system reaches its new energy equilibrium. An ideally linear systems homogenous response decays exponentially and never equals zero. In reality, the homogenous response of systems of practical interest eventually decays to
zero. We refer to the homogenous response as transient
because of its finite duration. Since an ideally linear homogenous response never equals zero, we must define the end
of the transient period. Conventionally, we set time equal to
either four or five times the systems time constant, , as
the duration of the transient period. The time constant scales
the rate exponential decay of the transient responses factors, and thus controls the rate at which a systems response
reaches steady-state.
The transient response of a system is the sum of the homogenous response and the particular solution for each input

Transient

Steady-State

x(t)

time, seconds
Fig. 2.11 An oscillatory step response showing the transient period
and the beginning of steady-state at time equal to five time constants

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 solution of the system equation represents the actual, observed
response of a dynamic system to the inputs acting on it.
Theresponse of a system is the superposition of the transient
response, created by disturbing the energetic equilibrium
between the system and the power supply, and steady-state
response to each input.

2.5 Method of Undetermined Coefficients


The Method of Undetermined Coefficients allows us to solve
linear ordinary differential equations with constant coefficients, by remembering a generic solution for the unforced
natural response, as well as generic solutions for a small number of physically common, forcing functions. The generic
solutions have adjustable constants (the undetermined coefficients) to scale the amplitude, duration, frequency, etc., to
fit the specific system equation, input, and initial conditions.
The general solution is constructed by the superposition, or
summation, of the natural and forced responses.
Our purpose in reviewing the method of undetermined
coefficients is to provide a rigorous justification for shortcuts
that we will develop. Our preferred solution method will be
to extract the information contained in the differential equation, in order to scale the known forms of the solutions to
match a specific system equation, and set of inputs. In short,
we will avoid as much of the mathematical formality typically found in a course in differential equations, as possible.
In that vein, consider the nth differential equation below:


dnx
d n 1 x
d n2 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
dt
dt
dt
dt

(2.10)

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

52
Table2.1 Common and
standard forms of ordinary
differential equations with
constant coefficients

Common form

Standard form

First-order
F (t ) = a 0

1
dv a1
F (t ) =
+ v
a0
dt a 0

dv
+ a1v
dt

a 0 dv
1
F (t ) =
+v
a1
a dt
1

Second-order
F (t ) = a 0

d 2v
dv
+ a1 + a 2 v
dt 2
dt

1
d 2 v a1 dv a 2
F (t ) = 2 +
+ v
a0
dt
a 0 dt a 0

d nv
d n1v
+ a1 n1 +  + a n v
n
dt
dt

an
1
d n v a1 d n1v
F (t ) = n +
++ v
a0
dt
a 0 dt n1
a0

nth-order
F (t ) = a 0

Although this nth order differential equation is in a general


form, it is also more difficult to read than a differential equation of a specific order. As we will see later, a fourth-order
differential system equation is a rarity of questionable accuracy. There is never a practical need for a higher-order equation. Consequently, we will illustrate the method of undetermined coefficients using a fourth-order differential equation,
so as to avoid the n notation.

Step 1. Separate the Input Functions and the Output Variables and Express the Differential Equation in Standard
Form.
The input functions, also known as forcing, exciting, or driving functions, describe what we plan to do, or did, to the system. Time is the independent variable in a dynamic system.
We must be able to describe our actions on the system in the
language of mathematics as functions of time. We will iden-


f1 (t ) + f 2 (t ) +  + f n (t ) = a 0

d4x
d3x
d2x
dx
+ a1 3 + a 2 2 + a 2
+ a4 x
4
dt
dt
dt
dt

Unfortunately, there are two different conventions for numbering the coefficients, each the reverse of the other. One
method assigns the subscript, zero, to the coefficient of the

tify the input functions as f (t ). The output variable, x, is also


a function of time. It should be written as x(t ), rather than as
simply x, but we normally dont 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



dt
dt
dt 
dt


Input Terms
Output Terms
highest-order term. The other method assigns the subscript,
zero, to the zero-order term. These conflicting conventions
are a problem whenever there is a formula expressed in
terms of the subscripts of the coefficients. Never use a formula which involves coefficient subscripts, unless you are
certain which numbering convention it is based on.
We will not limit the number of terms on the input side
of the differential equation. It is common and useful to
sum (or superpose) any number of inputs. Therefore, we
will retain the indefinite n for the number of input functions or terms.
The Method of Undetermined Coefficients is as follows:

(2.11)

(2.12)

On the left of Eq.2.12 is a summation of input functions.


This summation is the superposition of the input functions. The input summations occur for both physical and
mathematical reasons. Physically, it is generally difficult to
isolate a system from the surrounding environment. If the
action of the environment on the system is not negligible,
then there will be at least two physical inputs: what we do
to the system, and what the environment does to it. Multiple
input terms are created mathematically by different orders
of differentiation of the input function. Also, a single physical input may require the sum of two or more functions to
describe it. For example, a sinusoidal input with a non-zero
average value is the sum of a sinusoid and a Heaviside step

2.5 Method of Undetermined Coefficients

function. A second example is the construction of finite


duration pulses of various shapes, by summing scaled and
time shifted Heaviside step functions. If the differential
equation is linear and has constant coefficients, as described
below, then we can solve for the response of each input function separately, and sum (or superimpose) each of the results
to determine the overall response of the system.
We will use two different standard forms for differential
system equations, one form for first-order system equations,
and the other for second-order and higher-system equations.
The standard form for second-order and higher-system equations is as follows. After separating the input and output terms,
clear the highest-order output term of its coefficient, Eq.2.13.
This is the same standard form used with polynomials.

53

c. Is the equation linear? This takes more thought. The question of linearity refers to the dependent, or output, variable, x. The input or forcing functions can be non-linear.
In fact, an important forcing function is sin(w t ) . The
input side can have non-linear terms. It is the output side,
the side that describes the system, which must be linear.
The differential operator is a linear operator, regardless of
the order of the differential because it distributes onto a sum,
Eq.2.8., and doubling its input doubles its output, Eq.2.9.
For the output side to be linear, it cannot contain trig functions, powers of the dependent (output) variable, products of
the dependent variable and its derivatives, or products of the
derivative of the dependent variable and another derivative.
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
f1 (t ) +
f 2 (t ) +  +
f n (t ) = 4 +
x
+
+
+
a0
a0
a0
a 0 dt 3 a 0 dt 2 a 0 dt a 0
dt
The form we will use for first-order differential equations is
called time constant form. As we will develop in depth, the
time constant of a system, (Greek for t), is the time scale
of its natural or unforced response. The time constant can be
identified by units analysis, after the output variable term
of the first-order differential equation is cleared of its coefficient. The time constant is also known as the relaxation
time, a term with origins in elasticity, but now extended into
other applications.
Step 2. Check the Type of the Differential Equation.
The system equation produced by reducing a lumped parameter model of an energetic system will be a linear ordinary
differential equation, if the models of the individual energetic elements are all themselves linear models. In practice,
this means that equations which describe energy dissipation
in mechanical and fluid systems (i.e., friction), and some of
the equations which describe elastic strain energy storage
must be linearized, i.e., approximated by linear equations.
We will examine the process more closely in Chap.3, but the
need to eliminate non-linear element equations is absolute.
We cannot solve non-linear differential system equations
analytically. We must resort to computational (numerical)
methods, which we will develop in Chap.8. Check the system equation before proceeding.
a. Is it an ordinary differential equation? If there are partial
derivatives, then it is a partial differential equation, not an
ordinary differential equation.
b. Are the coefficients a 0 , a1 , a 2 , a n of the output side
constants, or are they functions of time? If they are not
constant, stop now! You cant solve a differential equation with non-constant coefficients, using the Method of
Undetermined Coefficients.

(2.13)

variable, v.


1 dF d 2 v b dv 2 K
= 2 +
+ v
M dt
M 
dt M
dt

(2.14)

Nonlinear

Most non-linear differential equations cannot be solved analytically, because you cannot use superposition to break the
problem down into bite-sized pieces. The solution of nonlinear differential equations is an all or nothing operation. Do
not attempt to use linear methods on non-linear equations.
Go back to the beginning and linearize the model, by using
linear element or constitutive equations.
Step 3. Check the Units of Each Term in the Equation.
This is an important check. You are checking to see if you
are adding apples and apples. All terms in a summation must
have the same units, or they cannot be summed. If the units
dont check, then you are working with nonsense. Check
now, before you invest any more time in the solution. Consider a second-order equation:
F = a0

d 2v
dv
+ a1 + a 2 v
2
dt
dt

where F is the input force in Newtons, and v is the output


velocity in m/sec.
What are the units of each term in this equation? Newtons, since the right side of the equation is force in newtons.
If this were a system equation, the coefficients, a0, a1, and a2
would consist of the parameters of the elemental (constitutive) equations. We will use SI units in all of our calculations.
Hence, the coefficients of mechanical system equations will

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

54

be terms with units of mass in kilograms, force in newtons,


displacement in meters, and time in seconds.
One should check units after expressing the equation in
standard form, since a unit check will reveal an error in that
process. Divide both sides by the coefficient of the highestorder derivative of the output variable.
F = a0

d 2v
dv
1
d 2 v a1 dv a 2
a
a
v
F
+
+

=
+
+ v
1
2
dt
a0
dt 2
dt 2 a 0 dt a 0

What are the units of the derivative terms? Recall the definition of a derivative from Calculus I.
dv
v
= lim
dt t 0 t
v
Therefore, dv has the units, , where the square brackdt
t
ets are read as units of. The derivative operator, d, means
infinitesimal change of. The derivative operator has no
units, and does not affect the units of the term it operates
2
on. Consider the second derivative, d 2v , and carefully note
dt
where the superscripts are. The superscript in the numerator
squares the derivative operator. The superscript in the de2
nominator squares the variable time, t. Consequently, d 2v
dt
v
has the units, 2 .
t

A fact of engineering mathematics you must know is the


solution of an ordinary differential equation with constant
coefficients homogeneous equation, Eq.2.16. This is the one
and only solution of the homogeneous equation:


0=

d 4 x a1 d 3 x a 2 d 2 x a 3 dx a 4
x
+
+
+
+
dt 4 a 0 dt 3 a 0 dt 2 a 0 dt a 0

(2.15)

This is called the homogenous equation, because homogeneous means self-generating, and the trivial solution of
the homogenous equation, x = 0, equals the input side of
zero.
The homogeneous equation represents the natural (unforced) response of the system, because the forcing function
is zero. Physically, the homogeneous solution represents the
response of the system, when it is given some amount of energy, then released. The energy flows through the system
and into the environment, in a way that depends on the nature 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
the system. Consequently, the homogeneous response of the
system is always present in the initial period of a systems
response. The duration of the homogenous response defines
the transient period. A system has reached steady-state,
when the homogeneous response has decayed to zero.

(2.16)

where n is the order of the differential equation. The coefficients An are the undetermined coefficients of the method of
undetermined coefficients. They are determined in the final
step of the method. It will be easy to remember Ae st because
we will use this exponential function frequently.
The complex variable, s = + jw , of the exponential Ae st is a characteristic value of the system. In other
words, it is a solution of the systems characteristic equation.
The characteristic equation is created by evaluating the homogenous equation with the trial solution, x = Ae st . Recall
how to differentiate an exponential.


deu
du
= eu
dx
dx

(2.17)

A source of confusion is how to differentiate the Laplace


variable, s, with respect to time. The Laplace variable s is
not a function of time. Consequently, it is treated as if it were
a constant, and its time derivative is zero.
dt
dAe st
de st
dst
ds
=A
= Ae st
= Ae st s
+t
dt
dt
dt
dt 0
dt 1

Step 4. Form and Solve the Homogeneous Equation.


Set the input side to zero (i.e., no forcing functions) to create
the homogeneous equation.


xH (t ) = A1e s1 t + A2 e s2 t +  + An e sn t

dAe st
= Ae st s
dt
Although the mnemonic from Calculus places du/dx after the
exponential, we will place the derivative in front, to make it
more prominent and, therefore, less likely to be lost during
transcription:


deu du u
dAe st
e
=
= sAe st
dx
dx
dt

(2.18)

Formulate and solve the homogeExample One


neous equation of the first-order differential equation,
dv a1 . Form the homogeneous equation by
1
a 0 F (t ) = dt + a 0 v
setting the input side of the equation equal to zero.
1
dv a1
dv a1
+ v 0=
+ v
F (t ) =
dt a 0
dt a 0
a0
The assumed form of homogeneous solution is always the
exponential, Ae st . Substitute the homogenous solution,
vH (t ) = Ae st , into the homogeneous equation:

2.5 Method of Undetermined Coefficients

0=

55

dvH a1
dAe st a1
Ae st
+ vH 0 =
+
dt
a0
dt
a0

Evaluate the derivative.


0 = sAe st +

a1
a0

Ae st

a1
0 = s + Ae st
a0

The parenthetical term is the characteristic function. For


this equation to be true, the characteristic function must
equal zero, since the exponential cannot. Setting the characteristic function equal to zero yields the characteristic equation, which is easily solved for s.
a1
a0

=0 s=

vH (t ) = Ae

(2.19)

Example TwoFormulate and solve the homogeneous


equation of the second-order differential equation,
1
d 2 v a1 dv a 2 .
F (t ) = 2 +
+ v
a0
a 0 dt a 0
dt
Form the homogeneous equation by setting the forcing
(input) function to zero.
d 2 vH a1 dvH a2
1
d 2 v a dv a2
+ v 0=
+
+ vH
F (t ) = 2 + 1
a0
a0 dt a0
a0 dt
a0
dt
dt 2

Substitute the homogenous solution, vH (t ) = Ae st, into the


homogeneous equation.

a1

0=

a0

The result is the homogeneous solution with the unknown


factor, A, the undetermined coefficient of the method of
undetermined coefficients.
st

d n Ae st
= s n Ae st
dt n

as repeated first-order differentiation. Each order of differentiation yields the Laplace variable, s, raised to the power
equal to the order of differentiation.

The exponential, e st , can never equal zero. Factor out the


exponential, Ae st.

s+

vH (t ) = Ae

a1
a0

d 2 Ae st a1 dAe st a 2
Ae st
+
+
a 0 dt
a0
dt 2

Evaluate these derivatives.


0 = s 2 Ae st +

a1
a0

sAe st +

a2
a0

Ae st

Physically, the natural response of all first-order, stable energetic systems is an exponential decay. We will see that homogenous, natural, or unforced, solution is present, even
when we drive a system with an input or forcing function.
The exponential decay results from the system reaching equilibrium to the changed conditions. If we have a system which
is initially de-energized, and apply a step input, then energy
will flow into the system, eventually leading to an equilibrium
condition, in which the amount of energy in the system stays
unchanged, until the input is again changed. Any subsequent
change to the input will initiate another natural response.
The homogenous solution of higher-order differential
equations follows the same steps as the previous first-order
system example. Higher-order derivatives of the solution
Ae st are evaluated by repeated differentiation,
d 2 Ae st
d dAe st d
dAe st
st
=
=
=
= s 2 Ae st
(
s
A
e
)
s
dt dt dt
dt
dt 2
where the Laplace variable s can be factored out of the derivative, since it is not a function of time. The shortcut is to
recognize that evaluating higher-order differentiation of Ae st
yields the factor, s, raised to the order of the differentiation,

Factor out the exponential Ae st .

a2
a1
0 = s 2 + s + Ae st
a0
a0

Set the characteristic function equal to zero, to form the characteristic equation.
s2 +

a1
a0

s+

a2
a0

=0

This algebraic equation is a second-order polynomial, in


which the Laplace variable, s, has two solutions, which are
found using the following quadratic equation:
s1 , s2 =

b b 2 4ac
2a
2

a1
a2
4
a0
a0
a0
a1

s2 +

a1
a0

s+

a2
a0

= 0 s1 , s2 =

Because the characteristic equation of the system has two


solutions, s1 and s2 , the homogenous solution is the sum of

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

56

two exponential terms of the form, Ae st. Either term solves


the homogeneous equation, but only the sum of both terms
creates the homogenous response of the system.
vH (t ) = A1e s1 t + A2 e s2 t
The two unknown values, A1 and A2, lead to the requirement
of two initial conditions to solve second-order differential
equations.
The significant difference between the homogeneous
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 higherorder equations may be complex numbers. Higher-order
differential system equations describe energetic systems
with two or more independent energy storage modes or

L {0} = L

d4x
4 +L
dt

a 1 d 3 x
+L

3
a 0 dt

d 4 x a 1 d 3 x a 2 d 2 x a 3 dx a 4
+
+
+
4 +
a 0 dt 3 a 0 dt 2 a 0 dt a 0
dt

L {0} = L

Although the Laplace transformation is only defined for


functions of time, zero is a special case, because of the linearity of the Laplace transformation. Constants can be factored
out of linear operators. Therefore, we could, if we choose,
multiply a function of time, f(t), by zero, and perform the
Laplace transformation on the product:

L {0 f (t )} = 0 L { f (t )} = 0 F ( s ) = 0
Distribute the Laplace transformation operator onto each
term of the summation.
a 2 d 2 x
+L

2
a 0 dt

a 3 dx

+L
a 0 dt

a 4

a 0

Factor out the coefficient terms.

L {0} = L

d 4 x a1
d3x a2
d 2x a3
dx a 4
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


is possible to store energy in it but no other energy storage
element in the system. The energy will not stay in that one
energy storage element. It will flow into the other elements
of the system. The energy storage element is independent,
if it can be energized for an instant when all other energy
storage elements are de-energized. A familiar example of
independent energy storage elements is a spring-mass system. We can visualize the spring extended or compressed
when the mass is stationary. A spring-mass system has two
independent energy storage modes or elements, strain or
potential energy and kinetic energy. When there are two
or more independent energy storage elements in a system,
energy can be transferred internally between energy storages within the system, as well as transferred into or out of
the system from the environment. Internal energy transfers
produce oscillations. Eventually, the oscillations die out,
as some of the energy during each internal transfer is dissipated as heat, and the system reaches equilibrium with
the input. Systems that have complex characteristic values
will oscillate, when subjected to a step input.
A more direct method of forming the characteristic equation is to perform the Laplace transformation on the homogeneous equation. The Laplace transform will be formally
introduced and its properties investigated in Sect.2.10.

The Laplace transformations of derivatives with respect to


time are Eqs.2.20 and 2.21.


df (t )

= sF ( s ) f ( 0)
dt

(2.20)

d 2 f (t )
df ( 0)
2

= s F ( s ) sf ( 0)
2
dt
dt

(2.21)

L
L

The initial condition terms, f (0) and df (0) , are neglected


dt
and only product of the transformed variable and the Laplace
variable s raised to the power of the order of the differentiation is retained.
Evaluate the Laplace transformation. An unknown timedomain variable, x(t), transforms to an unknown Laplacedomain variable X(s).
0 = s 4 X ( s) +

a1
a0

s3 X (s) +

a2
a0

s 2 X (s) +

a3
a0

sX ( s ) +

Factor out the Laplace-domain variable X(s).

a3
a2
a1
a
0 = s 4 + s3 + s 2 + s + 4 X (s)
a0
a0
a0
a0

a4
X (s)
a0

2.5 Method of Undetermined Coefficients

57

The summation is the characteristic function. For this


equation to be true, either the characteristic function equals
zero or X(s) equals zero. The latter is the trivial solution.
Setting the characteristic function equal to zero forms the
characteristic equation.
0 = s4 +

a1
a0

s3 +

a2
a0

s2 +

a3
a0

s+

a4
a0

Having recognized that the orders of s in the characteristic


equation equal the order of differentiation in the corresponding differential equation, we can omit the mathematical formality and write the characteristic equation directly. This
shortcut method can be dignified by calling it a mapping,
which is a systematic substitution.
d2
s2
dt 2

d
s
dt

dn
sn
dt n

Applying the shortcut mapping method:


0=

d 4 x a1 d 3 x a 2 d 2 x a 3 dx a 4
x
+
+
+
+
dt 4 a 0 dt 3 a 0 dt 2 a 0 dt a 0

a3
a2
a1
a
0 = s 4 + s3 + s 2 + s + 4 X (s)
a0
a0
a0
a0

0 = s4 +

a1
a0

s3 +

a2
a0

s2 +

a3
a0

s+

a4
a0

Characteristic equation is a good name, because the roots


of this equation (the ss that make it equal to zero) give the
time constant for the exponential decay and the frequency
of the oscillation, which indeed characterize the natural response of the system. A less meaningful but common name
for this equation is the eigen equation. While Euler was an
actual historical figure, there was no Herr Dr. Prof. Eigen.
Eigen is German for innate, or ones own. The roots of
the eigen (characteristic) equation are called the eigenvalues.
The term eigenvalue, a mixture of German and English,
is commonly used. An eigenvalue is a characteristic value.
The conventional notation for eigenvalues differs from that
of characteristic values. Lowercase lambda is used for an
eigenvalue, and lowercase s is used for a characteristic value.
Due to the likelihood of confusing the Laplace variable s
with the SI abbreviation s for seconds, we will depart from
standard SI notation and abbreviate seconds as sec.
The hardest part of solving higher-order ordinary differential equations with constant coefficients is solving the
characteristic equation. There are no general solutions for
polynomials above fifth order, and the general solutions for

third and fourth order are so time-consuming that, in the bad


old days of slide rules, we used iteration (guessing and refining guesses) to find the roots. In the twenty-first century, we
use software, such as Mathcad, MATLAB, and Mathematica, or high-end calculators, such as a TI-89 or TI-Nspire,
to solve polynomials. The software and calculators implement the JenkinsTraub algorithm to solve polynomials. The
algorithm estimates a factor of the polynomial, divides the
polynomial by that factor, and repeats the process until the
remainder is second order which is then solved by the quadratic formula.
Step 5. Solve the Forced Equation for Each Forcing Function.
There is a particular solution, xp, of the differential equation for each individual input or forcing function. The sum,
or superposition, of all of the particular solutions for the forcing functions is the forced response. The forced response is
the steady-state response of the system. All physical systems
try to follow their forcing functions. Linear systems actually
do follow theirs. The steady-state response of a linear system
will have the same functional form as the forcing function.
Therefore, find the particular solution for each individual
forcing function, by guessing a general solution that has the
same form as forcing function; substitute it into the differential equation; and evaluate the constants.
The trial solution should contain the function, and all of
its derivatives that produce a term that differs from the initial
term by more than a constant. Each term is preceded by an
unknown constant, hence the name, the method of undetermined coefficients. Common forcing functions include
steps, ramps, polynomials, sinusoids, and exponentials. Repeat Step 5 for each type of forcing function.
The most common forcing function is a step function of
some magnitude. For the time being, we will use step functions which turn on or transition at time, t = 0. A step
function is constant after it transitions. It is common to use
a constant instead of a step function, as a forcing function in
textbook problems. A ramp function is a step function multiplied by time, t. We will use, for the time being, ramp functions which transition on at time, t = 0. A parabola function
is a ramp function integrated with respect to time, t.
Summing (or superposing) a step and a ramp, creates a
polynomial forcing function:


f1 (t ) = K1 + K 2 t

for t > 0

(2.22)

Function f1 (t ) is a first-order polynomial, since t is raised to


the first power. Adding a parabola function makes the forcing function look more like a polynomial in t:


f 2 (t ) = K1 + K 2 t + K 3t 2

for t > 0

(2.23)

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

58

A trial solution, xp, must include each forcing function term


and its derivatives, if there are any. Considering the forcing function, f 2 (t ), the first term, K1, is constant and has no
derivatives.
x p 1 = C1
The second term, K 2 t, can be differentiated and expressed
as:
x p2 = C2 + C3t
The third term, K 3t 2, can also be differentiated. It and its
derivatives are
x p3 = C4 + C5t + C6 t 2
Superposing (or summing) the three trial particular solutions
x p1 + x p2 + x p3 = C1 + (C2 + C3t ) + (C 4 + C5t + C6 t 2 )
and collecting coefficients of like powers of time t yields the
following:
x p1 + x p2 + x p3 = C1 + C2 + C4 + (C3 + C5 )t + C6 t 2
Since the constants Cn are undetermined, terms of like powers of t can be combined, and the sum of the undetermined
coefficients represented by a single unknown, before substituting the particular solution into the differential equation.
Summing and renaming the unknown constants
C1 + C2 + C4 C4 and C3 + C5 C5
Hence, the particular solution for the polynomial forcing
function, f 2 (t ) ,


x p = C 4 + C5 t + C 6 t 2

(2.24)

The similarity between the mathematical forms of the input,


or forcing function, and the corresponding particular solution is not a coincidence. Physical systems to respond to their
inputs, and we will find that ideal linear systems replicate the
form of their forcing functions in their steady-state response.
A very important family of forcing functions is sinusoids.
Trial solutions for sinusoidal forcing functions require only
two terms, i.e.,


x p = C1 sin( t ) + C2 cos( t )

(2.25)

because of the relationship between the derivatives




d (C1 sin( t ))
= C1 cos( t ) = C1 cos( t )
dt

(2.26)

and


d (C2 cos( t ))
= C2 sin( t ) = C2 cos( t ) (2.27)
dt

since the coefficients C1 and C2 are undetermined, they can


absorb the angular frequency, , and the negative sign.
An important exception is the particular solution to be
used for exponential forcing functions. Exponential forcing functions are a special case, because exponentials are
also the solution to the homogeneous equation. If the forcing function is an exponential, f (t ) = Ae st , which, by rare
bad luck, happens to be exactly one of the exponential terms
in the homogenous solution, then the trial solution is modified, by multiplying it by time, t, so that instead of using
x p = Ae st , one uses this form:


x p = tAe st

(2.28)

Another special case occurs, if one root of the characteristic


equation is zero, and the system is subjected to a step input. If
s = 0, then Ae st = Ae s 0 = Ae0 = A, since any base raised to
the zero power equals one. Consequently, the constant term
in the homogenous solution has the same functional form of
one of the input terms, thestep input after it has transitioned
to a constant. The resolution to this problem is the same as in
the case of an exponential forcing function, where you may
multiplythe trial particular solution term by t. In this case,
the trial particular solution for a step input would be:


x p = Ct

(2.29)

The coefficients in the particular solutions are determined


by substituting the trial solution into the output side of the
equation, and collecting like powers of t or like trig functions.
Neatness is important to keep from drowning in constants.
There will be a factor comprised of the constants, an, from the
differential equation and the constants, Cn, from the particular
solution multiplying each term on the output variable side.
These factors must equal the coefficients of the corresponding
terms on the input side. If there is no corresponding term on
the input side, then the factor equals zero.
Solve for each of the coefficients, Cn. There are two reasons why one would never use initial conditions to find the
constants in a particular solution. First, the initial conditions
apply to the entire solution, not to just a part of it. A particular solution is part of the general solution. Second, the initial
conditions are irrelevant to a particular solution. A particular
solution describes the steady-state response of an energetic
system to a specific input. A system has reached steady-state
when the homogeneous solution has decayed to insignificance. The effect of the initial conditions decays with the homogeneous solution. An energetic system in steady-state is in

2.6 Initial Conditions

59

Table2.2 Initial conditions


needed for first, second, and
nth-order differential equations

Differential equation

Initial conditions needed

First-order

( )

1
dv a1
F (t ) =
+ v
a0
dt a 0

v 0+

a 0 dv
1
F (t ) =
+v
a1
a dt
1

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

( )

v 0+ and

( )

dv 0+
dt

nth-order
a
d n v a1 d n1v
1
F (t ) = n +
+ + n v
a0
dt
a 0 dt n1
a0

dynamic equilibrium with the power supplied by the forcing


functions. The steady-state response is solely dependent upon
maintaining the new dynamic equilibrium.
Step 6. Assemble the General Solution.
The general solution is the superposition (the sum) of the
homogeneous solution and the particular solution for each
forcing function, as seen here:


xh + x p = x

(2.30)

The general solution is the physically real solution. It is the


response that we see onthe oscilloscope. The initial conditions apply to the general solution! Do not apply the initial
conditions to the homogenous solution. The homogenous solution is not the complete solution.
Step 7. Use the Initial Conditions to Find the Undetermined
Coefficients.
The only undetermined coefficients in the general solution
are A1, A2, , An from the homogeneous solution. You will
have as many unknown coefficients from the homogeneous
equation, as you have orders of derivatives on the output
side of the system equation. You must have as many initial
conditions, as you have orders of derivatives; otherwise, you
cannot solve for the unknown coefficients. The initial conditions you need are those of the output variable and its n 1
derivatives. Determine the initial conditions using the energy
storage variables at time, t = 0 , and the energetic equations
used to derive the system equation. Solve for the unknown
coefficients by evaluating the general solution and its derivatives at time, t = 0+ .
Remember the mnemonic, Use the initial conditions
last.

( )

v 0+ ,

( ) , d v (0 )

dv 0+
dt

n 1

dt

n 1

Step 8.Check Your Solution Against Your Expectation of


the Response of the Physical System.
First, check the units in your solution. Are they consistent?
Then plot your solution. Does the solution agree with your
engineering judgment? In other words, does it seem to be
correct based on your understanding of the physical system?
If your solution doesnt make sense, then check it. Look for
the most common errors, such as sign errors and transcription errors first. If needed, rederive the system equation.
Never accept any computational result, until you believe it.
If you dont believe it, first look for errors. If there are no
mathematical errors, then the model is suspect. Reevaluate
your model. You may need to include energetic aspects of the
physical system that you initially believed were negligible.

2.6 Initial Conditions


Our purpose, when solving differential equations, will be
to predict the dynamic response of an energetic system. It
is impossible to predict the response of a system, if we do
not know its starting point. Hence, we need initial conditions. We must answer the following three questions: (1)
How many initial condition terms are needed? (2) What
initial conditions are needed? (3) How are those values determined?
The number of initial conditions needed equals the order
of the differential equation. For example, a first-order differential equation needs one initial condition, a second-order
differential equation needs two initial conditions, etc. We
must always know the value of the output variable at time,
t = 0+, which is the instant immediately after the input is applied. The additional initial condition(s), if needed, are the
values of the derivative(s) of the output variable at time,

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

60

t = 0+ . Using n as the order of the differential system equation, we must also know the values of n 1 derivatives of the
output variable at time, t = 0+.
We will address how values of the initial conditions are
determined in Chap.3, after we have discussed the energy
storage variables of an energetic system, and their role as
state variables. Briefly, if we wish to predict the response
of an energetic system to our input (a power source), we
must know how much energy is in each energy storage element in the energetic system, the instant before we apply
the input. We will see that knowing the input power variable (which we always know, since it is under our control)
and the values of the energy storage or state variables is
sufficient information to determine every other power variable in a system.

2.7 Complex Numbers and Variables


The coefficients of the method of undetermined coefficients
are functions of the characteristic values (eigenvalues) of the
system. The eigenvalues of second- and higher-order differential equations may be purely real numbers, complex conjugate pairs, or a combination of real numbers and complex
conjugate pairs. Eigenvalues which are complex conjugate
pairs yield a homogeneous solution which is a decaying sinusoid of the form:


f (t ) = Ae at sin (w t + )

(2.31)

Euler discovered the mathematics which relates complex


numbers and trigonometric functions. The exponential function with a purely imaginary exponent,


f ( jx) = e j x

(2.32)

has proved important in physics and engineering. It greatly


simplifies the representation and manipulation of decaying
oscillations. We will illustrate the utility of the complex exponential, e st = e( + jw )t = e t e jw t , and the use of Eulers sine
and cosine formula, in solving for the step response of an oscillatory second-order system, using the method of undetermined coefficients. We will first review complex numbers,
and then develop the mathematics of complex exponentials.

2.7.1 Cartesian Form Z = x + j y


The familiar Cartesian form of a complex number is as follows, Z = x + yi . It is conventional in systems dynamics
to use j rather than i as the symbol for 1, because i appears in our equations as electric current. We will view the

imaginary number, j, as the unit vector which defines the


imaginary direction of a complex plane. However, we will
not use conventional notation for a unit vector, and express
the imaginary number j as j for two reasons. First, as noted
above, complex numbers behave as vectors for somebut
not allvector operations. Specifically, complex numbers
are added as vectors; consequently, they can be decomposed
into real and imaginary components. However, we will see
that complex numbers do not multiply in the same manner
as vectors. Second, we would then also need to represent the
unit vector in the real direction. Unfortunately, the customary unit vector for the x-direction is i . If we were to use unit
vector notation, we would write the complex number as thus,
Z = xi + y j. This is unacceptable, since it is possible to confuse the unit vector in the positive real direction, i , with the
more common symbol for the imaginary number, i. Consequently, to avoid confusion, we will omit the unit vector in
the real direction. Also, to save time, we omit the vector notation for the imaginary number j. We will write the complex
number Z in Cartesian form as follows, Z = x + jy, with the
imaginary number j in front of the magnitude, y, to make it
more conspicuous.
The terms real, imaginary, and complex are unfortunate, since the latter two can intimidate students. Historically,
both real and complex numbers were discovered (or invented)
thousands of years after positive integers. Real numbers are
created by the arithmetic operations of addition, subtraction,
multiplication, and division. Complex numbers are created by
exponentiation of a negative real number to a fractional power
or by arithmetic operations on complex numbers alone or with
real numbers.
1

( 9) 2

9 = 3 1 = 3 j

An important aspect of complex numbers is the complex


conjugate. Conjugate means paired. Complex conjugates are
complex numbers with the same real component and equal
but opposite imaginary components as shown below.


Z, Z* = x jy

(2.33)

The superscript, *, which appears in Eq.2.33, indicates the


complex conjugate, as shown in Fig.2.12. Complex conjugates arise when solving polynomials. The roots of evenpowered polynomials with real coefficients can be complex
conjugate pairs. The common example is the quadratic formula. The roots are complex conjugates when the quantity
within the radial, the discriminant, is negative.
a1 a12 4a 2 a 0
 2
a 2 s + a1 s + a 0 = 0 s1 , s2 =
2a 2

(2.34)

2.7 Complex Numbers and Variables

61

Imaginary

Imaginary
jy

Z = x + jy

Z2 = 3+j

j3

Z 3 = Z1 + Z 2
= (-2+j2)+(3+j)
= 1+j3

j2

Z1 = -2+j2

x Real

-3

Z *= x - jy

-jy

Imaginary
j2

Z2 = 3+j

-3

-2

-1

-2

-1

3 Real

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

Z 1 = -2+j2

Z 2 = 3+j

3 Real

Fig. 2.13 The complex numbers, Z1 = 2 + j 2 and Z 2 = 3 + j

To use the parallelogram construction to perform subtraction, Z2 is multiplied by 1, which rotates the vector 180.
The rotated vector, Z2, is translated to the tip of Z1 to yield
the sum Z3, as shown in the vector construction, Fig.2.15.
There is a shortcut for constructing the difference between two vectors, which is to draw the resultant Z3 from
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
will make by using complex variables.

2.7.1.1Addition and Subtraction of Complex


Numbers
Addition and subtraction of complex numbers are easiest to
perform when complex numbers are expressed in Cartesian
form, as shown here:

2.7.1.2Magnitude (Modulus) of a Complex


Number
The mathematical term for the length or magnitude of a
complex number is its modulus. This is unfortunate, since
many other uses of the term modulus leads to confusion.
We will substitute the term magnitude instead. The Pythagorean Theorem yields the magnitude of the real vector
v, v = x1 + y1:

Z = Z1 + Z 2 = ( x1 + jy1 ) + ( x2 + jy2 )

= ( x1 + x2 ) + ( jy1 + jy2 )
= ( x1 + x2 ) + j ( y1 + y2 )

(2.35)

Z = Z1 Z 2 = ( x1 + jy1 ) ( x2 + jy2 )
= ( x1 x2 ) + ( jy1 jy2 )
= ( x1 x2 ) + j ( y1 y2 )

(2.36)

The graphical constructions, which are used to add or subtract position vectors or force vectors, are also used to add
or subtract complex numbers plotted as vectors in a complex plane. Complex numbers have magnitude, direction,
and obey vector rules for addition and subtraction, Fig.2.13.
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
complex numbers, as illustrated in the following examples.

v =

x12 + y12

(2.37)

The ancient Egyptians, Babylonians, and Greeks invented


(or discovered) geometry, long before either imaginary
numbers or negative numbers were invented. Consequently, Euclidian geometry and trigonometry use real numbers.
Physically realizable geometry, that is, what we can build, is
restricted to lengths, which are positive real numbers. Some
geometric relationships have been extended to include negative real numbers. Although the Pythagorean Theorem can
be used with negative real numbers, it cannot be used with
any imaginary numbers. The problem is the square of imaginary number


x12 + ( jy1 ) =
2

x12 + j 2 y12 =

x12 y12 Z1

(2.38)

Notice that the Pythagorean Theorem, as applied to


v = x1 + y1, yields the correct magnitude of Z1. Consequently,

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

62
Fig. 2.15 The difference between complex numbers, Z1 and
Z2, constructed using the parallelogram rule vector for vector
addition. Multiplying a vector or
complex number by 1 reverses
its direction

Imaginary
j2

-Z2 = -(3+j)

Z1 = -2+j2
j

Z 3 = Z1 - Z2
= (-2+j2)-(3+j)
= -5+j
-6

-5

-4

-3

-2

-1

-Z2 = -(3+j)

Z 2 = 3+j

3 Real

-j

Imaginary

Fig. 2.16 Vector shortcut to construct the difference, Z3=Z1Z2.


The difference Z3 is drawn from
the tip of Z2 to the tip of Z1

-Z2 = -(3+j)

Z 1= -2+j2

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

-5

Z 3= Z 1 - Z 2

-4

-3

-2

j2

-1

-Z2 = -(3+j)

Z 2 = 3+j
2

3 Real

-j

magnitude of a complex number is calculated using the Pythagorean Theorem, after removing the imaginary number

squared. We will consider the dot product of two complex


numbers to be undefined:

Z1 x1 + jy1 =

x12 + y12

(2.39)

2.7.1.3Multiplication and Division of Complex


Numbers
Although complex numbers add and subtract as vectors, they
are not, strictly speaking, vectors. Complex numbers both
lack operations of vectors and possess operations which vectors do not. The dot product of a real vector against itself
yields its magnitude squared.


v v = ( x1 + y1 ) ( x1 + y1 ) = x1 x1 + y1 y1

(2.40)

The dot product of a complex number against itself does not


yield the square of its magnitude.


Z1 i Z1 = ( x1 + jy1 )i( x1 + jy1 ) = x1 x1 + jy1 jy1


Z1 i Z1 = x1 x1 + j 2 y1 y1 = x1 x1 y1 y1

(2.41)

Consequently, the dot product of two complex numbers is either undefined, since it does not yield the magnitude squared,
or is defined as a special case, in order to yield the magnitude

Z1 Z 2 = undefined

(2.42)

Although the dot product of complex numbers can be defined for some applications, the cross-product of two complex numbers is not defined:


Z1 Z 2 = undefined

(2.43)

In one sense, we are splitting hairs, because there is no crossproduct defined for other quantities which we refer to as
vectors. For example, we may have a third-order dynamic
system with the state vector x comprised of the state variables, pressure, translational velocity, and torque. The crossproduct of the state vector, x, and a position vector, r, is not
defined either.
Complex numbers in Cartesian form are multiplied as an
algebraic expansion.


Z1Z 2 = ( x1 + jy1 )( x2 + jy2 ) = x1 ( x2 + jy2 ) + jy1 ( x2 + jy2 )


Z1Z 2 = x1 x2 + jx1 y2 + jy1 x2 + j jy1 y2
Z1Z 2 = x1 x2 y1 y2 + j ( x1 y2 + y1 x2 )

(2.44)

2.7 Complex Numbers and Variables

63

Im
Z1 = -2+j2

-1

j2
j

Z1

|Z 2|
1

Z2 = 3+j

|
|Z 1

|
|Z 1
-2

Z1 = -2+j2

Z2
3

Re

-2

Fig. 2.17 Complex numbers, Z1 and Z2, represented in polar form

A property which complex numbers share with scalars (but


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
complex conjugate of the denominator.


=
=

x1 ( x2 jy2 ) + jy1 ( x2 jy2 )

x1 x2 jx1 y2 + jx2 y1 j 2 y1 y2
x2 x2 jx2 y2 + jx2 y2 j 2 y2 y2
x1 x2 + y1 y2 + j ( x1 y2 + x2 y2 )

Re

Fig. 2.18 Complex numbers should be sketched before computing


their angles. Complex numbers falling in the second and third quadrant cannot be computed directly using the inverse tangent algorithm of
most calculators. A construction is needed

inverse tangent of the magnitude of the imaginary component over the real component.

Z2 in polar form is
(2.45)

There are two other forms for complex numbers. The first
is the polar representation, familiar from cylindrical coordinates, shown in Fig.2.17. Recall that a polar representation
of a vector in a two-dimensional plane specifies the length,
or magnitude, of the vector, and the angle that vector makes
with the positive x-axis. Likewise, the polar representation
of a complex number requires the length, or magnitude, of
the complex number, and the angle of the complex number.
The magnitude (modulus) is a purely real number. The angle
of a complex number is measured counter-clockwise from
the positive real axis. Unfortunately, the mathematical term
for the angle of a complex number is its argument, which
is more commonly used to identify the operand or input to
a function.
opposite
y
Z = tan 1
= tan 1
x
adjacent

Z 2 = 32 + 12 = 3.16

2.7.2 Polar Form: Z = Z  Z

Z1
j

The magnitude (modulus) of Z2 is calculated with the Pythagorean Theorem, using the magnitudes of the real and
imaginary components.

x2 ( x2 jy2 ) + jy2 ( x2 jy2 )

x22 + y22

j2

1
Z 2 = tan 1 = 0.32 rad
3

Z1
x + jy1
x + jy1 x2 jy2
= 1
= 1
Z 2 x2 + jy2 x2 + jy2 x2 jy2
=

-1

Im

(2.46)

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


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

Z 2 = ( Z 2 , Z 2 ) = ( 0.32 rad, 3.16)


The inverse tangent algorithm, commonly used by handheld
calculators, can introduce an error depending on the quadrant, in which the number complex falls.
The algorithm only returns angles between 90 and 90,
corresponding to complex numbers in the first and fourth
quadrants of a plane. To avoid error, always sketch a complex number as a vector, before calculating its angle. If the
complex number falls in the second or third quadrants, then
compute the angle using a geometric construction. For example, the angle of Z1 is calculated by subtracting angle
from 180 or rad, as seen in Fig.2.18 and Eq.2.47.
2
Z1 = 180o = 180o tan 1 = 180o 45o = 135o
2

1 2
 Z1 = = tan 2 = 4 = 2.36 rad
The magnitude of Z1 is
Z1 = 22 + 22 = 2.83
Z1 in polar form is
Z1 = ( Z1 , Z1 ) = ( 2.36 rad, 2.83)

(2.47)

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

2.7.3 Eulers Equations


Euler recognized that cos( ) + j sin( ) = e j . The proof of
the expression uses power series for cos(), sin(), and ex:

0 2 4 6
(2.48)
cos( ) =
+
+
0! 2! 4! 6!


(2.49)
sin( ) = + +
1! 3! 5! 7!
1

ex =

x 0 x1 x 2 x 3
+ + + +
0 ! 1! 2 ! 3!

(2.50)

e j = 1 + ( j )

( j )2 + ( j )3 + ( j )4 + ( j )5 + ( j )6 + ( j )7 
2!

e j = 1 + j +

3!

4!

5!

6!

7!

j 2 2 j 3 3 j 4 4 j 5 5 j 6 6 j 7 7
+
+
+
+
+

2!
3!
4!
5!
6!
7!

Use j 2 = 1,
e j = 1 + j

2
2!

j 3 4 j 5 6 j 7
+
+


3! 4! 5! 6! 7!

Collect terms with and without j.


2 4 6
j 3 j 5 j 7

+ j
+

e j = 1 +
2! 4! 6!
3!
5!
7!

Use the series for cos() and sin()




e j = cos ( ) + j sin ( )

(2.51)

e j

|=

j sin()

cos()

Re

Fig. 2.19 Eulers Equation is ej = cos() + jsin(). The projection of


ej onto the real axis is cos(). The projection onto the imaginary axis is
jsin(). The magnitude or length of ej is one

where e j is a complex exponential. We will refer to it as


a complex exponential unit vector, since its magnitude is
unity. Notice that the angle is multiplied by the imaginary
number j in the exponent. It is best to remember Eulers
equation, Eq.2.51, graphically, as shown in Fig.2.19, rather than symbolically.
The complex exponential unit vector is an important
quantity, as we shall see. The product of the complex exponential unit vector and a real number is a complex number
with the magnitude of the real number, and the direction or
angle of the complex exponential unit vector, Eq.2.52. This
is known as the complex exponential form of the complex
number.


By definition, 0 ! = 1 . Recall 0 = e0 = x 0 = 1.
Starting with the power series for ex, let x=j:

Im

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


to distinguish between an angle and the angles, = n 2
rad (or = n360o), since adding a complete revolution
about the origin of 2 rad (or 360) to the angle yields the same
orientation. The principal angle has two definitions. The principal angle of a complex number is either the smallest positive
angle, 0 < 2 rads ( 0 < 360o ), or the smallest angle,
rads ( 180o 180o ), needed to describe the
orientation of the complex number. The latter definition allows
a negative, purely real number to be described by either a positive or negative angle, which is useful in evaluating functions
which reach limiting values approaching the negative real axis
from either above or below.

|e

64

Z = Z e j

where = Z

(2.52)

Again, it is important to note that the angle of a complex


exponential unit vector is every term in the exponent which
is multiplied by the imaginary number j. Further discussion
follows in Sect.2.7.4.
The trigonometric projection of a point on a unit circle
centered on the origin onto the axes of a real plane is the
same as the projection of the complex exponential, e j , onto
theaxes of a complex plane, Fig. 2.20. The vertical axis of
the complex plane is the imaginary axis. Hence, the projection of e j onto the vertical axis is jsin().

2.7.3.1 Eulers Cosine Formula


Eulers cosine formula, Eq.2.53, is easiest to understand and
remember as a vector construction. The construction starts
with a complex exponential unit vector, e j . We will place it
in the first quadrant. The cosine formula is derived by summing e j with its complex conjugate, e j , which is its mirror
image on the other side of the real axis with the opposite
angle. Summing complex conjugates yields a purely real result, since the imaginary components are equal but opposite.
The vector construction, Fig.2.21, shows that the purely real

2.7 Complex Numbers and Variables

Im

Im
jsin()

-j

-e

2jsin()

|1
|

|e j
|=

sin()

65

cos()

Re

-j

-e

e j

jsin()

x 2 + y2 = 1

cos() + jsin() = e

|e

|=

cos()

Fig. 2.20a The relationship between the trigonometry of a unit


circle in a real plane and b Eulers formula in a complex plane,
e j = cos ( ) + j sin ( )

cos()

Re

e-j

Im

Fig. 2.22 Eulers sine formula sin ( ) =

e j e j
2j

|e

|=

jsin()

e -j

cos()

ed as separate factors. However, the complex exponential


form is a mathematical expression, whereas the polar form
is actually just the polar coordinates of the complex number. This distinction is illustrated below with the Cartesian
coordinates of a complex number versus the Cartesian form
of that number which is the sum of a real and an imaginary
number.

2cos() Re

-j

Fig. 2.21 Eulers cosine formula cos ( ) =

e j + e j
2

Z = ( x, jy )

result is twice the length of the projection cos( j). Consequently, the sum e j + e j must be divided by two.


cos ( ) =

e j + e j
2

(2.53)

2.7.3.2 Eulers Sine Formula


Eulers sine formula, Eq.2.54, is only slightly more complicated. Starting with e j , the complex exponential unit vector
is reflected relative to the real axis, by inverting the sign of
the exponent, i.e., the angle. The reflection is then rotated 180
by multiplying it by negative one, creating the vector, which
when summed with the original complex exponential, yields a
resultant which is purely imaginary and twice the length of the
projection, sin(), Fig.2.22. Consequently, it must be divided
by 2j to yield

e j + e j
sin( ) =
(2.54)
2j

2.7.4 Complex Exponential Form Z e j


The exponential form of a complex number Z e j resembles
the polar form, in that the magnitude and angle are represent-

Z = x + jy

A similar distinction exists between the polar coordinates of


a complex number and the complex number in exponential
form which is a product.
Z = ( Z, Z )

Z = Z e Z

The magnitude |Z| is calculated using the Pythagorean Theorem with the magnitudes of the real and imaginary components, Eq.2.39. Remember to drop j from the imaginary
component!
Z =

x2 + y 2

The angle is calculated using the inverse tangent function as


shown below:
opposite
y
Z = tan 1
= tan 1

x
adjacent
The algorithm, used in most calculators for the inverse tangent, only reports angles in the first and fourth quadrants of
the complex plane. Sketch complex numbers, before calculating the angle between them. This is demonstrated below
using the complex numbers Z4 and Z5 shown in Fig.2.23.

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

66

Z 4 = 1.41e j0.79

Im

Z 6= 2+j4 = 4.46 e j1.11

j4

|Z 5|

Z4

j3

Z5

Re

46
|

|Z

4|

Z 5 = 3.16e

j0.32

|4.

Im

j2

Z 6 =1.11

Fig. 2.23 Complex numbers, Z 4 and Z5 , in complex exponential form

Z4 = Z4 e

j Z 4

Z 4 = 1.41e

j 0.79

Z = 12 + 32 = 3.16
5

1 1
0
Z5 = tan = 36.9 = 0.32 rad
3

Z5 = Z5 e jZ5

Z5 = 3.16e j 0.32

Addition and subtraction of complex numbers expressed as


complex exponentials must be performed in vector form.
Consequently, addition and subtraction is easier in Cartesian
form than in complex exponential form, Fig.2.14. However,
multiplication, division, and exponentiation are much easier
in complex exponential form. Recall these three properties
of exponentials:

(2.55)
e a eb = e a + b



ea
= ea b
eb

(e )

a n

= e na

(2.56)
(2.57)

The properties of exponentials are independent of the constant we use as the base. We will illustrate the properties
using base 10, because its arithmetic is familiar, but the properties apply to base e, as well. Further, the properties apply
whether the exponents are constants, variables, or functions.
The product of exponentials equals the base raised to the
sum of the exponents.
102 103 = 102 + 3 = 105 and 105 = 101+ 4 = 101 104

4|

Z4

Z5 = 3.16e j0.32

|Z 5|
Z5

Re

Fig. 2.24 The product of complex numbers, Z 4 and Z5 in complex exponential form

and

Z5 = 3 + j

|Z

Z4 = 1 + j

Z = 12 + 12 = 1.41
4

1 1
0
Z 4 = tan = 45 = = 0.79 rad
1
4

Z 4 = 1.41e j0.79

The ratio of exponentials equals the base raised to the difference of the exponents.
102
105
23
1
3
5 2
10
10
and
10
10
=
=
=
=
103
102
The fact that any base raised to the zero power equals one is
easy to prove
10
= 10 1 10 1 = 10 11 = 100 = 1
10

2.7.4.1Multiplication in the Complex Exponential


Form
The product of two complex numbers in complex exponential form is evaluated by first associating the real magnitudes
and complex exponential unit vector factors.
Z 4 Z5 = Z 4 e j Z4 Z5 e j Z5 Z 4 Z5 = Z 4 Z5 e j Z4 e j Z5
The magnitude of the product Z4 Z5 is the product of the
magnitudes, |Z4| and |Z5|. The product of the complex exponential unit vectors is evaluated, using the property of exponentials expressed in Eq.2.55, Fig.2.24. The product of
exponentials equals an exponential raised to the sum of the
two exponents.
j Z + Z
e j Z4 e j Z5 = e j Z4 + j Z5 = e ( 4 5 )

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

Z 1 = 2.83e j2.36

67

Im
j2
j

Z1

Z 2 = 3.16e j0.32

Z2
-3

-2

-1

Z 7 = 4.24 rad

Z 7 = -2.04 rad

Re

Z 7 = 1.12e -j2.04
Fig. 2.25 The ratio of complex numbers Z2 over Z1, in complex exponential form

Transfer functions, introduced in Sect.2.10.4, are created


by performing the Laplace transform on a system equation,
and then creating the ratio of the transformed output variable
over the transformed input variable. The transformation of
the operation of differentiation, with respect to time, leads to
powers of the complex variable, s = + jw . A transfer function is thus a ratio of polynomials in the complex variable.


j Z + Z
Z 6 = Z 4 Z5 = Z 4 Z5 e ( 4 5 )

(2.58)

Z 6 = (1.41)(3.16)e j (0.79 + 0.32) = 4.46e j1.11

2.7.4.2 Division in the Complex Exponential Form


The ratio of two complex numbers in complex exponential
form is evaluated in a similar fashion. First, associate the
real magnitudes and complex exponential unit vector factors.
Z e j Z2
Z e
Z2
= 2 j Z1 = 2 j Z1
Z1
Z1 e
Z1 e

F (s) =

b0 s m + b1 s m 1 +  + bm
a 0 s n + a1 s n 1 +  + an

The ratio of the complex numbers Z1 over Z2 equals the ratio


of their magnitudes and the difference between the angles of
the numerator and denominator, Fig.2.25.


Z e j Z2
Z j
Z2
= 2 j Z1 = 2 e (Z2 Z1 )
Z1
Z1
Z1 e

Z7 =

(2.59)

Z 2 3.16e j 0.32 3.16 j (0.32 2.36)


e
=
=
= 1.12 e j 2.04
Z1 2.83e j 2.36 2.83

where m n (2.61)

F ( + jw ) =

Z N xN + jy N
=
Z D xD + jyD

(2.62)

This ratio must be reduced to a single complex number. One


method is to perform the division per Eq.2.45, i.e., to multiply numerator and denominator by the complex conjugate of
the denominator, which will yield a purely real number as the
denominator. Alternatively, the numerator and denominator
can both be expressed as complex exponentials, that is, by
using complex exponential unit vectors, and the ratio then
simplified, as shown in Eq.2.63.


e j Z2
j Z j Z
j
= e( 2 1 ) = e (Z2 Z1 )
e j Z1

(2.60)

The denominator of a transfer function is the characteristic


function which, when set equal to zero, forms the characteristic equation. Its roots are the characteristic values or eigenvalues of the system. Systems with complex conjugate
eigenvalues have oscillatory homogeneous responses.
When the numerator and denominator polynomials of the
complex function Eq.2.61 are evaluated for a specific complex value, such as s = + jw , the result is the ratio of two
complex numbers.

j Z 2

The magnitude of the ratio is the ratio of the magnitudes. The


ratio of the complex exponential unit vectors is evaluated
using the property of exponentials expressed in Eq.2.56.
The angle of the resultants complex exponential unit vector
equals the angle of the numerators angle minus the denominators angle.

N (s)
D (s)

In Eq.2.60, N(s) and D(s) represent the numerator and denominator polynomials. The general case is the following:


exponent of an exponential is the angle of the exponential.


The product Z4 Z5 is

F (s) =

F ( + jw ) = Z =

Z e j N
Z
ZN
= N j D = N e j ( N D )
ZD
ZD
ZD e

(2.63)

Unlike the roots of real numbers, the roots of complex exponentials are easy to calculate manually. We will illustrate
the properties of powers and roots of complex exponentials
using as an example the purely real number negative one expressed as a complex exponential, Fig.2.26. If the primary
angle is defined as the smallest angle which locates a complex number, then negative real numbers have two primary
angles, 180o = rad


1 = e j = e j

(2.64)

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

68

Im

e j = -1

Im

3 = = 3

e-j = -1

e = -1
-1

= -

j =e

= -1

j 2

1 = e j

Re

1 = e j 3

x = x , we see that the angle (argument) of the square root


of a complex number is one-half of the angle of the number,
Eq.2.65 and Fig.2.27.

=e

(2.65)

The imaginary axis is perpendicular to the real axis, thereby


confirming our calculation of the angle of j as one-half of the
angle of minus one which is 90 or /2 rad.
Notice that although we defined the imaginary number,
j, as the square root of negative one, there are actually two
square roots of negative one, j and j. Our preference for
positive numbers leads us to use j.
Another interesting example is the derivation of the cube
roots of negative one. If we express negative one as a complex exponential, and the cube root as a fractional exponent,
1 = e j = e j
3

1 = 3 e j = 3 e j

1 = e

( 1)3 = (e j )

1 = e j ( + 2 )

The last expression can be written as the product of two complex exponentials using Eq.2.55,

1
2

-j 2

=e

When the operation of obtaining the square root of a quantity


is expressed as the quantity raised to the one-half power, i.e.,

1 = e j = e j j = e

_
-j
3

=e

we find two of the three roots. The third root is negative one
itself. Negative one cubed equals negative one:

Fig. 2.27 Negative one has two square roots. We choose to use the
positive square root as the imaginary number, j

j2

Fig. 2.28 The cube roots of negative one

-
2 = __
2

-j = e

j2

1 = _
2

e j = -1
-j

j 1

Re

-
__
2 = 3

Fig. 2.26
Negative one expressed as a complex exponential,
1 = e j = e j

Im

= ej 3

1 = 3

Re

j1

=e

1 = e j ( + 2 ) 1 = e j e j 2
from which the following holds:


e j 2 = e j (2 2 ) = e j 0 = e0 = 1

(2.66)

Notice that negative one cubed is an alias of negative one.


Complex numbers expressed in polar or complex exponential form are not unique, meaning that there is not a single
complex number described by an angle and magnitude. It is
impossible to distinguish an angle from the angle 2n
since they overlie one another. An alias of a complex number
has the same magnitude (modulus) but with an angle 2n
rads greater or less than its principal angle.

2.7.5 Rotating Complex Exponential Unit Vector


We developed Eulers sine and cosine formula with a constant angle, . However, any quantity in the exponent of
an exponential multiplied by the imaginary number j is an
angle. An angle which is the product of time, t, and a constant frequency , (t ) = w t , where is the angular frequency in radians per second, increases at a constant rate. If
the angle has a non-zero value at time, t = 0, then the summation, (t ) = w t + is used, where the constant is the
angle at t = 0.

2.7 Complex Numbers and Variables

Im

69

1.0

e j(t+)

et

0.5

e j

et sin(t) = e-t sin(10t)

f1(t) 0

Re

-0.5

-et

Fig. 2.29 A rotating unit vector. The unit vector rotates in the positive,
counter-clockwise direction around the origin of the complex plane

A rotating complex exponential unit vector, which advances in the positive, counter-clockwise around the origin
of its complex plane as time increases, has a positive angular
frequency and a time-varying angle of either (t ) = w t or
(t ) = w t + , Fig.2.29.

-1.0

e j (wt + ) = cos(w t + ) + j sin(w t + )

(2.68)

A great utility of complex exponential unit vectors is the efficiency 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 et, and a sinusoid
with the form:
f1 (t ) = M 1e t sin (w t + 1 ) + C1

(2.69)

or


f 2 (t ) = M 2 e cos (w t + 2 ) + C2
t

(2.70)

where
f 2 (t ) =

M
2

t
e
cos (w t + 2 ) + C2





Amplitude Decay

Oscillation

Steady
State
Value

and w is the angular frequency in radians per second, is


the decay rate of the real exponential, oscillation, and is
the phase angle or phase shift. The term phase means
position in a cycle, as in phase of the moon.
An oscillation decays, Figs.2.30 and 2.31, if the real component of the eigenvalue, , is negative, since e at = 1at .
e
Conversely, the oscillation grows, Fig.2.32, if the real component of the eigenvalue, , is positive. It is an unstable exponential growth and has no limit in theory. An unstable system accumulates energy in the system at an increasing rate.
If the homogeneous response of the system is oscillatory, the

t, sec

Fig. 2.30 Plot of Eq. 2.69 with M 1 = 1, = 1 , w = 10 , 1 = 0 , and

C1 = 0

1.0

et

(2.67)
e jwt = cos(w t ) + j sin(w t )


Exponential Decay Envelope

0.5

Exponential Decay Envelope

= -1

et cos(t) = e-t cos(10t)

f 2(t) 0
-0.5

-et
-1.0

t, sec

Fig. 2.31 Plot of Eq. 2.70 with M 2 = 1, = 1, w = 10, 2 = 0, and


C2 = 0

150

= +1

100

Unstable Exponential
Growth Envelope et

et cos(t) = et cos(10t)

50

f 3(t) 0
-50
-100
-150

-et
0

t, sec

Fig. 2.32 Plot of Eq. 2.70 with M 2 = 1, = +1, w = 10, 2 = 0 , and


C2 = 0

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

accumulation of energy in the system increases the amplitude of the oscillations. In a physical system, eventually, if
the power supplied to the system is sufficiently large, the
amount of energy stored in the system exceeds the capacity
or strength of the system, thereby destroying it. If the power
supplied to the system is not large enough to destroy it, the
systems will reach a bounded unstable state, in which its
oscillations remain at the limiting amplitude. A bounded unstable state is a failure mode for the response of a system,
since we want systems to follow the input we give them, not
to go into endless oscillations.
The result of a dynamic calculation is presented in the
form of Eqs.2.73 or 2.74, as the product of a real exponential and a sinusoid. In fact, the execution of the calculation
is greatly simplified when the oscillation is represented by
a complex exponential unit vector instead of a sinusoid. We
will develop the mathematical technique below, and illustrate its use in example 2.7.6.

f(t) =

 f (t )] = Im[ Me st ] = Me t Im[e j wt ] = Me t sin(w t )


Im[

e st

b
1.0

j0.5

0.5

j0.0

0.0

-j0.5

-0.5

-j1.0
-1.0 -0.5

0.0

0.5

1.0

-1.0

Im[f(t)] = e tsin(t)

time

0.0

0.5

1
2
3
4

Fig. 2.33
a Plot of a rotating complex exponential,
f (t ) = Me st = Me( + jw )t , where M = 1 and s = 1 + j 2. b The projection

+
j
w
t
of Me( ) onto the vertical axis is Eq.2.68, Im f (t ) = e t sin (w t ) ,
c The projection of Me( + jw )t onto the horizontal axis is Eq.2.69,
Re f (t ) = e t cos (w t )

Im

j(t 2+)

j(t1+)

t 2
t1

-1

1 Re
-j

-t1
-j(t 2+)

-t2

-j

 f (t )] = Re[ Me st ] = Me t Re[e j wt ] = Me t cos(w t ) (2.73)


Re[
When the characteristic equation of a second-order differential system equation yields complex conjugate eigenvalues
(characteristic values), then the homogeneous solution will
be oscillatory. If input to the system is a step, then a solution (the response of the system) will exceed or overshoot
the final, steady-state value of the output variable. The functions, Re[ ] and Im[ ], are not as useful, when manipulating
complex conjugates, as Eulers cosine and sine formula,

1.0

(2.72)

Note that Im[ f (t )] is a purely real number. It is the magnitude of the imaginary component. The magnitude of the real
component is expressed as Eq.2.73.

Re[f(t)] = e tcos(t)
-1.0 -0.5
0

(2.71)
f (t ) = Me t e j w t = Me t + j w t = Me( + j w ) t = Me st
This expression is a rotating exponential unit vector, with a
time varying magnitude. If the real component, , is negative, the vectors magnitude decreases with time. The tip of
the vector traces a logarithmic spiral, Fig. 2.33a, a familiar
natural shape. A cross-section of a snails shell is a logarithmic spiral, because a snails growth rate is proportional to its
size. The trace of a snails growth is a spiral from the origin
out. Conversely, the decay rate of a dynamic systems oscillation is proportional to the amount of energy stored in the
system. The trace of the response of a dynamic system is a
spiral toward the origin. Software and calculators have the
functions Re [ ] and Im [ ], which return only the magnitude
of the real or imaginary components of a complex number,
respectively. The imaginary component of the complex exponential is expressed, as Eq.2.72,

e te jt = e (+j)t

j1.0

time

70

-j(t 1+)

Fig. 2.34 Counter-rotating complex conjugate exponential unit vectors, e j(wt + ) and e j(wt + )

Eqs.2.49 and 2.50, reproduced below. The corresponding


vector constructions are Figs.2.22 and 2.23.
cos ( ) =

e j + e j
2

and

sin ( ) =

e j e j
j2

2.7 Complex Numbers and Variables

71

Eulers cosine and sine formulas contain complex exponential unit vectors with equal but opposite angles. When the
angle is the time-varying angle = w t + then the complex exponential unit vectors e j (wt + ) and e j (wt + ) counter-rotate about the origin of a complex plane, Figs.2.33 and 2.34.
The resultant of e j (wt + ) + e j (wt + ) lies on the real axis,

with constant coefficients which have complex conjugate eigenvalues and, hence, an oscillatory, homogeneous solution,
with the following example:

(2.74)
e j (wt + ) + e j (wt + ) = 2 cos (w t + )

These are the initial conditions: v(0+ ) = 0 and

or

Likewise, the resultant, e j (wt + ) e j (wt + ) , drawn from the tip


of e j (wt + ) to the tip of e j (wt + ), is parallel to the imaginary
axis:

The units are not consistent in this textbook example. The


input is a constant, rather than a function of time, but we will
ignore that physical nonsense, and proceed to the solution.
We will begin by solving the homogeneous equation. The
first step is to formulate and solve the characteristic equation. Again, the solution of the homogeneous equation is always the following:

e j (wt + ) e j (wt + ) = j 2sin (w t + )

vh (t ) = A1e s1t + A2 e s2t

e j (wt + ) + e j (wt + )
= cos (w t + )
2

(2.75)

(2.76)

or


e j (wt + ) e j (wt + )
= sin (w t + )
j2

(2.77)

The complex exponential unit vector may be intimidating at


first, because it is new mathematics. However, it is much
easier to deal with than the corresponding product of a real
exponential, and the sum of the product cosine and sine, to
yield a sinusoid with phase shift, using one of the four following trigonometric identities:


sin ( + ) = sin ( ) cos ( ) + cos ( ) sin ( )

(2.78)

sin ( ) = sin ( ) cos ( ) cos ( ) sin ( )

(2.79)

cos ( + ) = cos ( ) cos ( ) sin ( ) sin ( ) (2.80)

cos ( ) = cos ( ) cos ( ) + sin ( ) sin ( ) (2.81)

The expansion of the expression from the complex exponential form to the trigonometric form, with the phase shift
represented as the sum of products of sinusoids, is reason
enough to work with complex exponentials:
e t (e j (w t + ) + e j (w t + ) ) = 2e t cos(wt + )
= 2e t (cos(wt ) cos( ) sin(wt ) sin( ))

2.7.6Example: Solution of a Homogeneous


Equation using Complex Exponential
Unit Vectors
We will illustrate the use of complex exponential unit vectors to solve a second-order ordinary differential equation,

8=

d 2 v dv
+ + 6v
dt 2 dt
dv(0+ )
= 10.
dt

The number of terms equals the number of eigenvalues


which equals the order of the differential equation. Forming
the characteristic function by means of the Laplace transformation

L {0} = L
L {0} = L

d 2 v dv

2 + + 6v
dt
dt

d 2v
dv
2 + L + L {6v}
dt
dt

0 = s 2V ( s ) + sV ( s ) + 6V ( s ) 0 = s 2 + s + 6 V ( s )
Assuming V(s)=0 is the trivial solution. Set the characteristic function equal to zero, to form the characteristic equation.
s2 + s + 6 = 0
The roots of the characteristic equation are the characteristic
values or eigenvalues
s1 , s2 = jw

s1 , s2 =

1 12 4 6
2

s1 , s2 = 0.5 j 2.40
It is more efficient and less error-prone to perform as much
of the solution symbolically as possible. Substitute in the numerical values of the eigenvalues only when necessary.
We now solve the particular equation, using the form of
the input as the form of the particular solution. The input
is a constant. Hence, the particular solution is the unknown
constant C, as follows:
vp = C

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

72

Substitute and evaluate as follows:


8=

d 2vp
dt 2

dv p

+ 6v p

dt

C=

8=

dC
d 2C
+
+ 6C
2
dt 0 dt 0

Now, evaluate dv(t ) for the known value at time, t = 0+ :


dt

( ) = 10 = A s e

dv 0+

1 1

dt

8
= 1.33
6

We have two equations with two unknowns, A1 and A2.

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

1.33 = A1 + A2

+
We now use the initial conditions, v(0 ) = 0 and
+
dv(0 ) / dt = 10, last with the general solution, to determine
the undetermined coefficients, A1 and A2:

( )

v 0+ = 0 = A1e s1 0 + A 2 e s2 0 + 1.33
0 = A1 e0 1 + A 2 e0 1 + 1.33
0 = A1 + A2 + 1.33 1.33 = A1 + A2

10 = A1 s1 + A2 s1
Resist the temptation to substitute the values of the eigenvalues into the equations. Numbers are only needed for arithmetic.
Algebra is easier with symbols. It is easier to see the required
operations and requires less writing to execute them. Choose a
solution method from: (1) direct substitution and elimination,
(2) Gaussian elimination, or (3) linear algebra. Direct substitution is easy for a system of equations with two unknowns.
1.33 = A1 + A2 A1 = 1.33 A2

In order to use the initial condition, dv(0 ) / dt = 10, differentiate the general solution and then evaluate it for time,
t = 0+ . Reversing this order and evaluating the general solution, before differentiating it, will mistakenly lead to identifying quantities as constants, which are, in fact, variables
that happen to have a known value at the instant, t = 0+.
dA 2 e
dv (t ) dA1e
=
+
dt
dt
dt
s1t

dt

dA1e
dt

+ A2

s2 t

dA1e

dC
dt

dt

= A1

de s1t
de s2t
+ A2
dt
dt

dt

dv (t )
dt

dt
dt
+ A 2 e s2t s2
dt
dt

= A1 s1e s1t + A 2 s2 e s2t

10 = 1.33s1 + A 2 ( s2 s1 )

10 + 1.33s1
= A2
s2 s1
10 + 1.33s1
s2 s1

Now, substitute for s1 and s2 symbolically as s1 , s2 = jw .


A2 =

10 + 1.33s1
s2 s1
A2 =

de
du
= eu
.
dx
dx

The eigenvalues, s1 and s2, are constants. Factor them out of


the derivatives:
= A1e s1t s1

10 = 1.33s1 A 2 s1 + A 2 s2

dv (t )
ds t
ds t
= A1e s1t 1 + A2 e s2t 2
dt
dt
dt

dv (t )

A1 = 1.33 A2 A1 = 1.33

dt

Recall how to differentiate an exponential

10 = A1 s1 + A2 s2 10 = 1.33 A2 s1 + A2 s2

s2 t

Factor the undetermined constants, A1 and A2, out of the derivatives


dv (t )

10 = A1 s1 + A2 s2

v (t ) = vh (t ) + v p (t ) v (t ) = A1e s1t + A2 e s2t + C

dv (t )

+ A 2 s2 e s2 0

10 = A1 s1 e0 1 + A 2 s2 e0 1

Assemble the general solution:

s1t

s1 0+

A2 =

10 + 1.33 ( + jw )

jw ( + jw )

10 + 1.33 + j1.33w
j 2w

Express A2 as a single complex number rather than as a ratio.


j 2w
Multiply by the unity ratio,
.
j 2w
A2 =
A2 =

10 + 1.33 + j1.33w j 2w
j 2w
j 2w
j ( 20w + 2.66w ) + j 2 2.66w 2
j 2 4w 2

2.7 Complex Numbers and Variables

A2 =
A2 =

73

j ( 20w + 2.66w ) 2.66w 2


4w

2.66w 2
20w 2.66w
+ j
+

4w 2
4w 2
4w 2

5 + 0.67
A 2 = 0.67 j

w
A1 = 1.33 A 2

A1 = 1.33 0.67 +

5 + 0.67
j

5 + 0.67
A1 = 0.66 j

w
Comparing A1 and A2, we see that they are complex conjugates. The discrepancy between real components is a rounding error, due to limited numerical precision. We now substitute in the values, = 0.5 and w = 2.96 .
5 + 0.67
A1 = 0.66 j

Having determined the undetermined coefficients and expressed them as complex exponentials, we now substitute
them and eigenvalues into the general solution symbolically.
v (t ) = A1e s1t + A 2 e s2t + 1.33
v (t ) = Me j e( + jw )t + Me j e( jw )t + 1.33
Distribute time t in the exponents.
v (t ) = Me j e t + jw t + Me j e t jw t + 1.33
Using Eq.2.55, we express the complex exponentials as the
products of a real exponential and a complex exponential
unit vector.
v (t ) = Me j e t e jw t + Me j e t e jw t + 1.33
Factor out the product Me t.

v (t ) = Me t e j e jw t + e j e jw t + 1.33
Now use Eq.2.55 to combine the products of two exponential.

5 + 0.67 ( 0.5)
A1 = 0.66 j

2.40

v (t ) = Me t e j + jw t + e j 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 0.67
A 2 = 0.67 + j

v (t ) = Me t e jw t j + e jw t + j + 1.33

A 2 = 0.67 + j1.94

v (t ) = Me t e j (w t ) + e j ( w t + ) + 1.33

Express A1 and A2 as complex exponentials. Make a simple


sketch of the complex numbers to ensure that correct angles
(arguments) are calculated, since A1 and A2 are in the third
and second quadrants of the complex plane, respectively.

The exponents are equal but of opposite sign, as needed to


use Eulers sine or cosine formula. It is easier to recognize
that the exponents are of opposite sign, which reduces the
risk of a sign error with the phase shift, , if the negative sign
of the exponent of the second exponential is factored out.

1.94
A2 = A1 = tan 1
= 1.90 rad
0.67
The magnitude (modulus) of the complex A1 and A2 is the
following:
A1 = A2 = 0.67 2 + 1.942 = 2.05
thus,

A1 = 2.05e j1.90 and A2 = 2.05e j1.90

The angle of A1 is the phase shift of the response sinusoid.


We will express A1 and A2 as complex exponentials:
A1 Me j and A2 = Me j

v (t ) = Me t e j (w t ) + e j (w t ) + 1.33
Note that the complex exponentials are counter-rotating
unit vectors with the phase angle, -. The sum of the counter-rotating complex exponential unit vectors is in the form
of Eq.2.74. After we multiply the first term by the unity
ratio of 2/2.
v (t ) =

2 t j (w t + )
Me e
+ e j (w t + ) + 1.33
2

v (t ) = 2 Me t

e j (w t + ) + e j (w t + )
+ 1.33
2

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

74

6
-0.5t

v(t) = 4.10 e

cos(2.40t-1.90)+1.33

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


Step 8. Check your solution against your expectation of the
response of the physical system.

4.10 e-0.5t +1.33


2

2.8.1Example Problem One: First-Order


Step Response

v(t)
0

-4.10 e-0.5t +1.33

Solve the following first-order ordinary differential equation


with constant coefficients:

-2
-4

6
0

time

Fig. 2.35 Plot of the solution of 8 =

( )

ditions, v 0+ = 0 and

( ) = 10

dv 0+

10

d v dv
+ + 6v with the initial condt 2 dt

dt

We use Eulers cosine formula to obtain the form of Eq.2.70:


e j (w t + ) + e j (w t + )
v (t ) = 2 Me
+ 1.33
2

We need the initial conditions to calculate the response. In


physical systems, we must know, directly or indirectly, how
much energy is stored in the systems energy storage elements, immediately prior to the application of the input. It is
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
instant after we apply the following input:

( )

v 0+ = 0

v (t ) = 2 Me cos (w t + ) + 1.33
t

We now, finally, substitute for and from s = + jw


= 0.5 + j 2.40 and for M and from Me j = 2.05e j1.90:
v (t ) = 2 ( 2.05) e 0.5t cos ( 2.40t + ( 1.90)) + 1.33
v (t ) = 4.10e 0.5t cos ( 2.40t 1.90) + 1.33
The result is plotted in Fig.2.35.

2.8Solved Problems Illustrating the Method


of Undetermined Coefficients
We shall illustrate the Method of Undetermined Coefficients
by applying it in three examples.
Method of Undetermined Coefficients
Step 1. Separate the input functions and the output variable.
Express the equation in standard form.
Step 2. Check the nature of the differential equation.
Step 3. Check the units of each term in the equation.
Step 4. Solve the homogeneous equation.
Step 5. Solve the forced equation for each input function to
yield the particular solutions.
Step 6. Assemble the general solution by summing (superposing) the homogeneous and particular solutions.

dv
+ 3v = 8
dt

where t = 0+ is the instant after the input is applied at t = 0.


Step 1.Separate the Forcing Functions and the Output
Variable. Express the equation in standard form.
The output variable is v. The independent variable is t. The
input (or forcing function), 8, is nonsense. Again, a forcing function (the input) must be a function of time; it cannot be a constant. We will presume that the input is 8us (t ),
where the Heaviside unit step function us (t ) transitions from
zero to one, when its argument becomes positive, at the time
t we define as t = 0+ . Express the differential equation in
standard form, by clearing the coefficient from the highest
derivative of the output variable.
6

dv
+ 3v = 8us (t )
dt

dv 3
8
+ v = us (t )
dt 6
6

dv
+ 0.50 v = 1.33us (t )
dt
Step 2. Check Units of Each Term in the Equation.
dv
dt + [ 0.50v ] = 1.33us (t )
dv
dt + 0.50 [ v ] = 1.33 us (t )
v
t + [v] = [ ]

2.8 Solved Problems Illustrating the Method of Undetermined Coefficients

The units are inconsistent. This is not a valid equation. The


units of each term in the summation and on both sides of the
equation must be identical, or we are equating apples with
oranges. We will ignore that, for this example. In a system
equation, the coefficients carry units from the energy storage
and dissipation element equations, yielding a valid equation.
Step 3. Check the Nature of the Differential Equation.
dv
+ 0.50 v = 1.33us (t )
dt
This is an ordinary differential equation, since we are only
differentiating with respect to one variable, time=t. The coefficients of each term are constants. Hence, the differential
equation is an ordinary differential equation with constant
coefficients. We can use the Method of Undetermined Coefficients to solve it.
Step 4. Solve the Homogeneous Equation.
The homogeneous equation is formed by setting the input
side of the equation to zero.
dv
+ 0.50v = 0
dt
We solve the homogeneous equation by knowing the form
of the solution. The homogeneous solution is always of the
form Ae st . The number of terms equals the order of the differential equation. The homogeneous solution of this firstorder differential equation has one term.
vh (t ) = Ae st
The characteristic value or eigenvalue, s, is the solution of
the characteristic equation. There are two methods and one
shortcut for forming the characteristic equation from the homogeneous equation:
1. Substituting the homogeneous solution, vh (t ) = Ae st , into
the homogeneous equation yields the characteristic equation of the system.
2. An alternative method to the characteristic equation is to
perform the Laplace transformation, neglecting the initial
condition terms.
3. The shortcut is a mapping or substitution, based on the
Laplace transformation, in which each time derivative is
replaced with the corresponding power of s.
d
s
dt

d2
s2
dt 2

dn
sn
dt n

75

To illustrate, we will first use substitution of the homogeneous solution, vh (t ) = Ae st, followed by differentiation:
dv
+ 0.50v = 0
dt

dvh (t )
dt

+ 0.50vh (t ) = 0

dAe st
+ 0.50 Ae st = 0
dt
Recall how to differentiate an exponential,
Ae s t

deu
du
= eu .
dx
dx

d ( st )
+ 0.50 Ae s t = 0
dt

sdt tds
st
+
Ae s t
+ 0.50 Ae = 0
dt
dt
The variable s is not a function of time t. Hence,
td ( s )
= t 0 = 0
dt
and
sdt tds
st
st
st
+
Ae st
+ 0.50 Ae = Ae ( s1 + t 0) + 0.50 Ae
dt
dt
( s + 0.50) 
Ae st = 0 s + 0.50 = 0 s = 0.50
Never
Zero

Now, the alternative technique of mapping d s


dt
dv
+ 0.50v = 0 sv + 0.50v = 0 ( s + 0.50)v = 0
dt
The trivial solution, ( s + 0.50)v = 0, is v = 0. The trivial
solution is what gives homogeneous solution its name, homogeneous or self-generating, since 0 = 0. The characteristic equation is a first-order polynomial in the variable, s.
It has one root, s = 0.5, which is the characteristic value, or
eigenvalue, of the system.
( s + 0.50)v = 0 s + 0.50 = 0 s = 0.50
The homogeneous solution carries the unknown, yet-to-bedetermined coefficient, A:
vh (t ) = Ae st = Ae 0.5t
Step 5. Solve the Forced Equation for Each Forcing Function.
The trial solution for each input is a summation. The first
term of the trial solution has the same functional form as the

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

76

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


first term. The input in this problem is a Heaviside step function, Eq.2.1. A true Heaviside step change, in which a variable has a value at time, t = 0 , and a different value at time,
t = 0+ , can never be obtained in an actual physical system,
above the atomic level of quantum mechanics. However, we
closely approximate step changes in many circumstances.
Consequently, we will find the Heaviside step function extremely useful mathematically. The derivative of a step, after
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
step takes place over zero time, or defined as the impulse
singularity function, (t ), which also cannot exist physically
above the atomic level. All physical approximations of impulses are pulses of finite duration. However, we will also
find the impulse function useful mathematically.
We will omit the impulse from the forced (or particular)
solution, because our solution is only valid, after the input
has been applied to the system, i.e., for time, t > 0+ . The impulse input serves to initialize the homogeneous response.
The homogeneous response is initiated by any change of the
input to the system and, hence, is also present in the step
response. We would keep the impulse input it were the only
input term, but otherwise it is not needed. Consequently, the
forced (or particular) solution is the same form as the step
input:
v p (t ) = C us (t ) = C

for

t > 0+

where C is a constant.
Substitute and evaluate derivatives.
dv p (t )
dt

+ 0.50v p (t ) = 1.33

0.50C = 1.33

dC
+ 0.50C = 1.33
dt 0
C = 2.66

Step 6. Assemble the General Solution.


The general solution is the superposition (sum) of the following homogeneous and particular solutions
v(t ) = vh (t ) + v p (t )
v(t ) = Ae st + C
v(t ) = Ae 0.5t + 2.66
Step 7. Use the Initial Conditions to Find the Undetermined
Coefficients.
The initial conditions allow us to evaluate the general solution for time t = 0+ . Any base raised to the power of zero
equals one, a 0 = b 0 = 1. Hence,

v(t)

v(t) = 2.66-2.66 e-0.5t

vh (t) 0
vp (t)

-0.5t

-2

-4

v h(t) = -2.66 e
0

time

10

Fig. 2.36 Homogeneous, particular, and general solution of Example


dv
+ 3v = 8
dt

One, 6

e( 0.5)( 0 ) = e0 = 1
0 = A + 2.66 A = 2.66
The general solution follows and is plotted in Fig.2.36:
v(t ) = vh (t ) + v p (t ) = 2.66e 0.5t + 2.66
The expression is a stable exponential growth function. It is
stable because the eigenvalue, the multiplicative constant in
the exponent, is negative. The function approaches zero, as
time approaches infinity. If the exponent were positive, then
the response would be unstable, and the function would approach infinity, as time approached infinity.
We will rewrite general solutions as,
v(t ) = 2.66(1 e 0.5t )
This form shows the amplitude (or magnitude) and the nature
of the response more clearly. The multiplicative constant is
the magnitude of the response. You will soon recognize (or
associate) the nature of the response by the function.
v (t ) = 2.66


(1 e )
0.5t



Magnitude 
Stable Exponential
Growth Function

Note that the constants are expressed as decimal fractions.


You convey the confidence you have in a result by the number of significant figures you report. Nothing in engineering
is exact. In fact, one objective of this text is to introduce
some of the many approximations necessary in engineering
modeling. Carry as many significant figures in decimals as
you wish as intermediate results during a calculation, but
never report results to more than three significant figures,

2.8 Solved Problems Illustrating the Method of Undetermined Coefficients

77

unless measurements of all of the parameters, the input variable, and the initial value of the output variable were made to
unusually high precision, and you have confidence that the
model is accurate to that precision.

d 2 v dv
2 + 3 + [1.5v ] = [ 4us (t ) ]
dt dt
v
v
t 2 + 3 t + 1.5 [ v ] = 4us (t ) [

2.8.2Example Problem Two: Non-Oscillatory


Second-Order Step Response
Solve the following second-order ordinary differential equation with constant coefficients:
2

d dv

+ 3v = 8us (t ) 3v
dt dt

v v
t 2 + t + [v] = [ ]
These units do not match, but we will ignore this, as it is a
textbook problem.
Step 3. Check the Nature of the Differential Equation.
d 2v
dv
+ 3 + 1.5v = 4us (t )
dt
dt 2

dv(0+ )
= 0.
where the initial conditions are v(0 ) = 0, and
dt
+

Step 1.Separate the Forcing Functions and the Output


Variable. Express the differential equation in standard form.
First, distribute the derivative operator, d :
dt

This is an ordinary differential equation with constant coefficients, and we can use the Method of Undetermined Coefficients to solve it.

d 2v
dv
d dv

+ 3v = 2 2 + 3
dt
dt dt
dt

Step 4. Solve the Homogeneous Equation.


Form the homogeneous equation by setting the input side of
the equation to zero.

d 2v
dv
2 2 + 3 = 8us (t ) 3v
dt
dt
Now, separate the forcing function and the output variable:
d 2v
dv
2 2 + 3 + 3v = 8us (t )
dt
dt
Finally, write the differential equation in standard form,
where the highest-order derivative of the output variable has
no coefficient:
d 2v
dv
+ 3 + 1.5v = 4us (t )
2
dt
dt
Step 2. Check Units of Each Term in the Equation.
d 2 v dv
2 + 3 + [1.5v] = [4us (t )]
dt dt
Recall

v
dv
dv
v
= lim . Hence dt has units t . The units

dt t 0 t

d 2v
2 are less apparent. They are easier to identify in this
dt
d 2 v d dv v
form: 2 = = 2 .
dt dt dt t

d 2v
dv
+ 3 + 1.5v = 0
dt
dt 2
Since the differential equation is of second order, the homogeneous solution will have two terms.
vh (t ) = A1e s1 t + A2 e s2 t
Form the characteristic equation, using the method of your
choice. Performing the Laplace transformation (neglecting
the initial condition terms) yields:

d 2v
dv
2 + L 3 + L {1.5v} = L {0}
dt
dt
s 2V ( s ) + 3sV ( s ) + 1.5V ( s ) = 0

( s 2 + 3s + 1.5)V ( s ) = 0 s 2 + 3s + 1.5 = 0
Solve the second-order polynomial in s. Second-order polynomials are solved using the quadratic equation:
s1 , s2 =

3 (3) 2 (4)(1.5)
2

s1 = 0.63

and

s2 = 2.37

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

78

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


evaluated using t = 0+ :

yielding
vh (t ) = A1e 0.63t + A2 e 2.37 t

dv (t )
dt

We will use the initial conditions last to determine A1 and A2.


Step 5. Solve the Forced Equation for Each Forcing Function.
d 2v
dv
+ 3 + 1.5v = 4us (t )
dt
dt 2
Cause and effect requires that the response of the system follows the application of the input. Therefore, we need only
evaluate the forcing function for the condition, t 0+ . The
Heaviside unit step function transitions from zero to one,
when its argument becomes positive.
us (t ) = 1 for
d 2 v p (t )
dt

+3

dv p (t )
dt

C=

4
= 2.67
1.5

Step 6. Assemble the General Solution.


Superimpose (sum) the homogeneous and particular solutions:

v (t ) = A1e 0.63t + A2 e 2.37 t + 2.67


Step 7. Use the Initial Conditions to Find the Undetermined
Coefficients.
We must have as many initial conditions as the order of the
differential equation, because we will have that many undetermined coefficients. Each initial condition yields an equation, and we need as many equations as unknowns.
dv(0+ )
The initial conditions are v(0+ ) = 0 and
= 0. Set
dt
v(0+ ) = 0,

( )

( 0.63)(0+ )

+ A2 e

( 2.37)(0+ )

Again, recall that exponentials are differentiated as


deu
du
= eu :
dx
dx
dv (t )
d ( 2.37t )
d ( 0.63t )
= A1e 0.63t
+ A2 e 2.37 t
dt
dt
dt
dv (t )
d (t )
d (t )
= A1e 0.63t ( 0.63)
+ A2 e 2.37 t ( 2.37 )
dt
dt
dt

Evaluate for t = 0+ :

( ) = 0.63 A

dv 0+

dt

+ 2.67 = 0

A1 + A 2 = 2.67
We must differentiate the general solution to use the initial condition, dv(0+ ) / dt = 0. A common error is to reverse
the order of the operations expressed in dv(0+ ) / dt = 0. The

( 0.63)(0+ )

2.37 A 2 e

( 2.37)(0+ )

=0

0.63 A1 2.37 A 2 = 0
We now have two equations with two unknowns:
A1 +A2 = 2.67
0.63 A1 2.37 A2 = 0

v(t ) = vh (t ) + v p (t )

v 0+ = A1 e

dv (t )
= 0.63 A1e 0.63t 2.37 A2 e 2.37 t
dt

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


d C
dC
+ 1.5 C = 4
+3
dt 0
dt 2 0

d
d
d 2.67
A1e 0.63t +
A 2 e 2.37 t +
dt
dt
dt 0

dv (t )
d 2.37 t
d 0.63t
= A1
e
+ A2
e
dt
dt
dt

t>0

+ 1.5v p (t ) = 4

We can solve this system of equations with a variety of techniques. Direct elimination is straightforward. Eliminate A2
by rearranging the first equation, and substituting the result
into the second:
A2 = 2.67 A1

0.63 A1 2.37 2.67 A1 = 0


Solve for A1,
A1 =

(2.37)(2.67)
= 3.64
(2.37 0.63)

Back substitute for A2


A2 = 2.67 A1 = 2.67 ( 3.64) = 0.97
The general solution is
v (t ) = vh 1 (t ) + vh 2 (t ) + v p (t )

2.8 Solved Problems Illustrating the Method of Undetermined Coefficients

79

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


d 2v
dv
2 + 0.33 + [ 0.5v ] = [1.33 us (t ) ]
dt
dt

v(t) 2
vh (t)
vh (t)

vp (t)

v v
t 2 + t + [ v ] = [

v(t) = -3.64 e -0.63t + 0.97 e-2.37t+ 2.67

-2.37t

vh (t) = -3.64 e

The units do not check. If this were a system equation, we


would not proceed, until we eliminated the inconsistency in
the units. You can multiply apples and oranges, but you cannot add them.

-2

-0.63t

vh (t) = -3.64 e
1

-4

time

10

Fig. 2.37 Homogeneous, particular, and general solutions of Example


d 2v
dv
Two, 2 + 3 + 1.5v = 4
dt
dt

Step 3. Check the Nature of the Differential Equation.


It is an ordinary differential equation with constant coefficients. We can use the Method of Undetermined Coefficients.
Step 4. Solve the Homogeneous Equation.
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


and is plotted in Fig.2.37.

2.8.3Example Problem Three: Oscillatory


Second-Order Step Response

d 2v
dv
+ 0.33 + 0.5v = 0
dt
dt 2
d
Create the characteristic equation by mapping
s and
dt
factoring out the output variable:
s 2 v + 0.33sv + 0.5v = 0

Solve the following second-order ordinary differential equation with constant coefficients:

dv(0+ )
= 0.
dt

Step 1. Separate the Inputs (Forcing Functions) and Output


Variable.
The forcing function and the output variable terms are already separated. We will express the differential equation in
standard form by clearing the highest-order derivative of its
coefficient:
d 2 v 2 dv 3
8
+
+ v = us (t )
6
dt 2 6 dt 6
d 2v
dv
+ 0.33 + 0.5v = 1.33 us (t )
2
dt
dt

(s

+ 0.33s + 0.5 v = 0

s 2 + 0.33s + 0.5 = 0
Solve the characteristic equation:

d 2v
dv
+ 2 + 3v = 8us (t )
2
dt
dt

+
where the initial conditions are v(0 ) = 0 and

s1 , s2 =

s1 = 0.16 + j 0.68
0.33 (0.33) 2 (4)(0.5)

s2 = 0.16 j 0.68
2

Note that the characteristic equation has complex conjugate


roots. Complex conjugate roots identify an underdamped
system which will oscillate, when disturbed by any input.
The imaginary component of the root is the frequency of the
oscillation in radians per second. The real component of the
root is the decay rate of the oscillation.
There are two forms of the homogeneous solution. The
first is the product of a real exponential and the sum of a sine
and a cosine term:
vh (t ) = e 0.16t ( A1 cos(0.68t ) + A2 sin(0.68t ))

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

80

The second form of the homogeneous solution uses the complex exponential unit vector introduced in Sect.2.7.4:
st

Using

s t

dv (t ) d 0.16 t
A1 cos ( 0.68 t ) + A 2 sin ( 0.68 t ) + 2.66
=
e
dt
dt

vh (t ) = A1e 1 + A2 e 2 = A1e( 0.16 + j 0.68)t + A2 e( 0.16 j 0.68)t

The complex exponential form is easier to work with, and


will be our preferred method. However, we will demonstrate
both methods, starting with the trigonometric homogeneous
solution.

1.33
= 2.66
0.5

( )) 0.68e(

dv(0+ )
= 0:
dt

( ) sin 0.68 0+
( )( )

0.16) 0+

+ A 2 0.16e

0 = A1 0.16 e0 1 cos ( 0) 0.68 e0 1 sin ( 0)


1

Now, evaluate for t = 0+ using the initial condition

Step 7. Use the Initial Conditions to Find the Undetermined


Coefficients.
cos ( 0.68) 0+

v(t ) = e 0.16t ( A1 cos(0.68t ) + A2 sin(0.68t )) + 2.66

( 0.16)(0+ )

dv (t )
= A1 0.16e 0.16 t cos ( 0.68 t ) 0.68e 0.16 t sin ( 0.68 t )
dt
+ A 2 0.16e 0.16 t sin ( 0.68 t ) + 0.68e 0.16 t cos ( 0.68 t )

v(t ) = vh (t ) + v p (t )

dv(t )
d
d
= A1 (e 0.16t cos(0.68t )) + A2 (e 0.16t sin(0.68t ))
dt
dt
dt

Step 6. Assemble the General Solution.

0 = A1 0.16e

dv (t ) d 0.16 t
d 0.16 t
A 1 cos ( 0.68 t ) +
A 2 sin ( 0.68 t )
=
e
e
dt
dt
dt
d
+ ( 2.66)
dt
0

The forcing function 1.33us (t ) = 1.33 for t > 0. Hence, the


particular solution is a constant, C.
C=

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 )
2
dt
dt

The differential operator must be distributed on the right side.


Although not necessary, before we do so, we will distribute
the exponential onto the cosine and sine terms to make the
subsequent differentiation slightly easier.

Step 5. Solve the Forced Equation for Each Forcing Function.

d 2C
dC
+ 0.5C = 1.33
+ 0.33
dt 0
dt 2 0

dv(0+ )
=0
dt

))

( 0.16)(0+ )

) + A ( 0.16 e
2

( ))

sin 0.68 0+ + 0.68e


0

( 0.16)(0+ )

sin ( 0) + 0.68 e0 1 cos ( 0)


0

( )))

cos 0.68 0+

0 = 0.16 A1 + 0.68 A2

dv(0+ )
The initial conditions are v(0 ) = 0 and
= 0 . Using
dt
v(0+ ) = 0
+

0=e

( 0.16 )( 0+ )

( A1 cos((0.68)(0 )) + A2 sin((0.68)(0 ))) + 2.66

0 = e 0 ( A1 cos(0) + A2 sin(0)) + 2.66

0 = e 0 1 A1 cos ( 0) + A 2 sin ( 0)
1

0 = A1 + 2.66

+ 2.66

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


yields the coefficients A1 and A2:
A1 = 2.66
0 = A1 + 2.66
A1 = 2.66

0.16
0 = 0.16 A1 + 0.68 A 2
A 2 = 0.63
A2 =
A1
0.68
Substituting the now determined coefficients, A1 and A2,
into the general solution yields the result which is plotted in
Fig.2.38:
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

f1(t)

f3(t) = f1(t)+f2(t)

81

f 2 (t) = -0.63 sin(0.68t)

1
0.5

M x = -0.63

-2 -1.5

-1 -0.5

0.5

1.5

-0.5

f2(t) 0

-1

-1.5

f3(t)

-4

-2

-2

-2.5

M y = -2.66
-3

f1(t) = -2.66 cos(0.68t)


0

10

time

15

20

Fig. 2.38 Plots of f1 (t ) = 2.66cos ( 0.68t ), f 2 (t ) = 0.63sin ( 0.68t ) ,


and the sum f3 (t ) = f1 (t ) + f 2 (t ). The sum of sinusoids is difficult to
visualize

A drawback to the trigonometric form of the homogeneous


solution is that the resulting waveform, the sum of a sine and
a cosine term, is difficult to visualize. We know the amplitude of the individual sine and cosine terms, but what is the
amplitude of the sum? Further, what is the phase shift of the
sum?
The sum of a cosine and a sine term with the same frequency and, therefore, the same angle at any time t, can be
expressed as single sinusoid at that frequency, plus a phase
shift . The calculation of the phase shift is based on one of
the following four trigonometric identities, Eqs.2.78 to 2.81.
We will use the sine relationship, Eq.2.78.
sin ( + ) = sin ( ) cos ( ) + cos ( ) sin ( )
Equating the identity, Eq.2.74, with the trigonometric factor
of our result:
sin ( ) cos ( ) + cos ( ) sin ( )

= 2.66 cos ( 0.68 t ) 0.63sin ( 0.68 t )

We see that to equate with t and with , we must rearrange the terms:
sin (w t + ) = sin ( 0.68 t )( 0.63) + cos ( 0.68 t )( 2.66)
When we attempt to equate the factors terms, cos() and
sin(), with the corresponding constants, we discover a
problem:
cos ( ) = 0.63 and sin ( ) 2.66

Fig. 2.39M is a vector sum, M = 0.63x 2.66y

The limits of sine and cosine are negative one and one. We
must insert a scaling factor M with an unknown magnitude
before cos() and sin() in the identity, thus,
 M sin (w t + )

= sin ( 0.68 t ) M cos ( ) + cos ( 0.68 t ) M sin ( ) (2.82)

which yields:
M cos( ) = 0.63and M sin( ) = 2.66
We can interpret this pair of equations as the projection of
a vector M onto the horizontal and vertical axes of a real
plane:
M = M cos ( ) x + M sin ( ) y

M = 0.63x 2.66y

Note x and y denote unit vectors in the positive x and y directions, to avoid the potential confusion that the conventional
notation i and j could cause. We will plot the vector, before
we attempt to find its angle . We find that the vector M lies
in the third quadrant, Fig.2.39.
The angle is calculated indirectly, as the difference between and the angle between M and the negative x-axis:
2.66
0.63

= + tan 1

= + 1.34

= 1.80 rad
We calculate the magnitude of M via Pythagorean Theorem,
thus,
M = M x2 + M y2

M =

M = 2.73

( 0.63)2 + ( 2.66)2

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

82

5.5
-0.16t

v(t) = 2.73e

sin(0.68t-1.80)+2.66

Repeating Step 7. Use the Initial Conditions to Find the


Undetermined Coefficients.
The initial conditions are v(0+ ) = 0 and dv(0+ ) / dt = 0 .
Using v(0+ ) = 0 :

2.73e-0.16t+2.66

v(t)
2

v(0+ ) = A1e

-2.73e-0.16t +2.66

10

time

work with the variables, s1 and s2, as symbols rather than


values whenever possible. It saves time and reduces the possibility of a transcription or sign error.

( s1 )( 0+ )

+ A2 e

( s2 )( 0+ )

+ 2.66 = 0 A1 + A2 = 2.66

Using dv(0+ ) / dt = 0:
20

dv (t )

30

Fig. 2.40 General solution of Example Problem Three


d 2v
dv
6 2 + 2 + 3v = 8us (t ) with the initial conditions, v 0+ = 0 and
dt
dt

( )

( )=0

dv 0+

dt

d
d
d
s t
s t
A1e 1 +
A 2 e 2 + ( 2.66)
dt
dt
dt
0

dv(0+ )
( s )(0+ )
( s )(0+ )
= s1 A1e 1
+ s2 A2 e 2
= 0 s1 A1 + s2 A2 = 0
dt
We have a system of two equations with two unknowns:

dt

A1 + A 2 = 2.66
We check these results by evaluating M cos ( ) = 0.63 and
M sin ( ) = 2.66 for M = 2.73 and = 1.80 rad:
2.73 cos( 1.80) = 0.62 and 2.73 sin( 1.80) = 2.66
Substituting M and into Eq.2.78:

s1 A1 + s2 A 2 = 0
As illustrated in Sect.2.7.5, we will not substitute in numerical values for s1 and s2 until necessary. Solve for A1 and A2.
Eliminate A2:
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)

(s

We can now express the general solution in terms of sine


with the phase angle , Fig. 2.40:
v(t ) = 2.73e 0.16t sin(0.68t 1.80) + 2.66
We now illustrate the alternative solution, in which the oscillation of the homogeneous solution is represented by complex exponentials. We return to Step 6, and now use this expression,
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

s2 A1 = s2 2.66 A1 =

A2 = A1 2.66 A2 =

A1 =

2.66 s2
s1 s2

s t

A1 =

v (t ) = A1e(0.16 + j 0.68)t + A 2 e(0.16 j 0.68)t + 2.66


Note how much more cumbersome is the general solution,
when the values of s1 and s2 are substituted in. It is best to

s1 s2

2.66 s2
s1 s2

2.66

Substitute for s1 and s2 symbolically, where s1 = + jw and


s2 = jw

v (t ) = A1e 1 + A 2 e 2 + 2.66
s t

2.66 s2

Solve for A2:

Repeating Step 6. Assemble the General Solution.


v (t ) = vh (t ) + v p (t )

s1 A1 + s2 A1 2.66 = 0 s1 A1 s2 A1 = s2 2.66

A2 =

2.66 s2
s1 s2

A1 =

2.66 ( jw )
+ jw ( jw )

2.66 ( jw )
j 2w

2.66 A2 =

2.66 ( jw )
2.66
+ jw ( jw )

2.8 Solved Problems Illustrating the Method of Undetermined Coefficients

A2 =

2.66 ( jw )
2.6
j 2w

A1 =

A1 =

A1 =

A1 =

2.66 ( j ( w ))
2w

j 2.66 2.66w
+
2w
2w

2.66 j j 2w
j 2 2w

1.33

2.66 ( jw ) j
j 2.6
j 2w

A2 =

A2 =
A2 = j

2.66 j j 2w

2.66 ( j ( w ))
2w

j 2w
2

2.66 2.66 w
+
2.66
2w
2w

1.33

Re
-j0.32

A 1 = -1.33 - j0.31

0.32
1.33

2 = tan 1

2 = 0.24 = 2.91 rad

2.66 ( j + w )
2w

A 2 = 1.33 + j

2.6

A2 = 1.33 + j

Complex conjugates have equal magnitudes (moduli)


and equal but opposite angles (arguments), Fig.2.41.
Substitute in numerical values for and where
s1 , s2 = jw = 0.16 j 0.68,
1.33 ( 0.16)
= 1.33 j 0.31
0.68

A2 = 1.33 + j

1.33 ( 0.16)
= 1.33 + j 0.31
0.68

0.32
1.33

1 = + tan 1

2 = + 0.24 = 2.91 rad

1.33

A1 = 1.33 j

1 = +

1.33

Comparing A1 and A2, we see they are complex conjugates:


A1 = 1.33 j

-1.32

2 =

) 2.6

2.6 A2 =

j0.32

Calculate the magnitude (modulus) of the complex conjugate coefficients, A1 and A2, by the Pythagorean Theorem.
Angles 1 and 2 should be calculated indirectly, to avoid the
error most calculators make in evaluating the inverse tangent
of vectors in the second and third quadrants. Calculate angle
using the inverse tangent and positive real values for the
lengths of the adjacent and opposite sides of the right triangle, which includes :

Continuing with A2:


A2 =

Fig. 2.41 Complex conjugate coefficients, A1 and A2.

2.66 ( j + w )
2w

A1 = 1.33 j

Im

A2 = -1.33 + j0.31

A1 and A2 contain ratios of a complex number over an imaginary number. Both coefficients should be expressed as a
single complex number. Multiply by the unity ratio j/j to
eliminate the imaginary number in the denominator. Starting
with A1:
2.66 ( jw ) j
A1 =
j
j 2w

83

A1 = A2 =

( 1.33)2 + (0.31)2

= 1.37

Express A1 and A2 as complex exponentials:


A1 = A1 e

j A1

= 1.37e j 2.91 and A2 = A2 e

j A2

= 1.37e j 2.91

Substitute into the homogeneous solution the complex exponential form of A1 and A2:
vh (t ) = A1e( 0.16 + j 0.68)t + A2 e( 0.16 j 0.68)t
vh (t ) = 1.37e j 2.91e( 0.16 + j 0.68)t + 1.37e j 2.91e( 0.16 j 0.68)t
Distribute the time variable, t, and then use the property of
exponentials, e a + b = e a eb , Eq.2.51, to express the exponentials with exponents that are sums, as the product of exponentials:
vh (t ) = 1.37e j 2.91e 0.16t + j 0.68t + 1.37e j 2.91e 0.16t j 0.68t

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

84

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


Factor out the constant and the purely real exponential:

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

Use the property, e a + b = e a eb , to express the product of two


exponentials as a single exponential:

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

Factor the imaginary number j out of the exponents:

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

We will rearrange frequency and phase angle terms in the exponents into the conventional order, w t + , where w = 0.68
and = 2.91:

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

Recall that any quantity multiplied by j in the exponent of an


exponential is its angle. The angles are equal but of opposite
sign. Factoring out the negative sign from the second exponential makes this easier to see:

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

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


is compared with Eulers cosine and sine formulas:
cos( ) =

e j + e j
Eq. 2.53
2

sin( ) =

e j e j
Eq. 2.54
j2

we see that the cosine formula can be used, if the exponential


term is multiplied by the unity ratio 2/2:
vh (t ) = 1.37e 0.16t
vh (t ) = 2.74e 0.16t

2 j (0.68t 2.91)
e
+ e j (0.68t 2.91)
2

e j (0.68t 2.91) + e j (0.68t 2.91)


2

vh (t ) = 2.74e 0.16t cos ( 0.68t 2.91)


Assemble the general solution:
v (t ) = vh (t ) + v p (t )
v (t ) = 2.74e

0.16 t

cos ( 0.68t 2.91) + 2.66

This result differs from our previous by a rounding error of


0.01.

2.9Eigenvalues and Response


Characterization
As we have seen, the homogeneous solution of an ordinary
differential equation with constant coefficients is the sum
of exponentials of the form Aest, where the variables, s, are
the roots of a differential equations characteristic equation.
These roots are the differential equations characteristic values or eigenvalues. The eigenvalues are the coefficients of
the variable time t in the exponentials of the homogeneous
solution. They determine the temporal or time scaling of the
homogeneous solution. Eigenvalues are either purely real
numbers, e.g., s = , or complex conjugates, s1 , s2 = jw .
Complex conjugate eigenvalues indicate that the homogeneous response is underdamped and oscillatory. Oscillatory
homogeneous responses have two time scales, the period of
the oscillation and the decay rate of the oscillation.
The time scaling of step responses reveal the eigenvalues of the system. Consequently, the time scaling of a step
response obtained from a test of an existing physical system
is used to develop a mathematical model of the system. This
process is known as system identification. A second use
of the eigenvalues time scaling is to establish the duration
of the response. Further, when plotting a response function,
eigenvalues time scaling can be used to establish the time
step, which is the time interval between successive calculations.

2.9.1 First-Order Step Responses


A purely real eigenvalue, or the real component of a complex
eigenvalue in an exponent, yields a real exponential which,
as a factor, scales the amplitude of the output variable. We
have seen that a real eigenvalue or the real component of a
complex eigenvalue must be negative for the response function to decrease or decay with time and be stable. Conversely, if a real eigenvalue or the real component of a complex conjugate eigenvalue is positive, the real exponential
formed by the eigenvalue grows with time, and the response
is unstable.
The characteristic time of a real exponential is the inverse
of the absolute value of a purely real eigenvalue or the real
component of a complex eigenvalue, and is known as the
time constant. The time constants symbol is tau, , Greek
for t, Eq.2.83:


(2.83)

The magnitude or absolute value removes the negative sign


of the eigenvalue or its real component, of stable systems.

2.9 Eigenvalues and Response Characterization


Fig. 2.42 a Unit exponential
decay with time scaled in
units of time constants .
b Unit stable exponential
growth with time axis scaled
in time constants . The
responses decay 63.2% the
remaining difference from
their current and steadystate values during each
time interval, equal to one
time constant

85

1
63.2%

_t
e-

86.5%

98.2% 99.3%

_t
1- e-

86.5%

95.0% 98.2% 99.3%

63.2%

A negative sign must be added to the exponent, when the


exponential is written in terms of the time constant.
t

95.0%

time

f 2 (t ) = f 2 1 e 2t + C2

time

f 2 ( t ) = f 2 1 e 2 + C2

(2.84)

(2.87)

A property of exponents implied by Eq.2.56 is that a negative exponent expresses an inverse fractional relationship:

The dynamic range f is the absolute value of the difference


between the initial and final values of the step response. We
will examine the step responses of first-order systems in detail in Chap.3.
The real exponential of a first-order systems step response is characterized by its time constant . A simple
method to determine the time constant from data is to calculate either 63.2% of the change in the output variable from
its initial to final value, or the corresponding 36.8% of the
change remaining, and then to find the time corresponding
to that value in the data, interpolating if necessary. Estimating a time constant from the duration of a first-order step
transient is usually inaccurate, because noise in the data and
the quantification imparted by digital instrumentation tend
to obscure when the signal reaches steady-state defined as
either 98.2%, 4, 99.3%, or 5, of the step change.

e t = e

for

e a =

<0

1
ea

(2.85)
t



1
t

If a = , then e = t . Evaluating e at the limits


e
of t = 0 and t = yields

=
e

1 1
= = 1 for t = 0
e0 1

and
e

=
e

1
1
= = 0 for t =
e

Since e decreases from one to zero as time proceeds from


zero to infinity, it is a decay function. Although a decay
function requires infinite time to reach zero, it approaches

Example OneDetermine the response function which


describes the data plotted in Fig.2.43.
The response is a decay in the form of Eq.2.86 with
C1=0, since the response decays to zero. The dynamic range
is the initial value minus the steady-state value.

f1 = f1 ( 0+ ) f1 ( ss )

zero quickly. Evaluating e


for time expressed in multiples of the time constant , Fig.2.42 serves as illustration:
A unit exponential decay, Fig.2.42a, and a unit stable exponential growth, Fig.2.42b, range from zero to one. In a
step response, such as that shown in Fig.2.36, these functions are scaled in amplitude by a coefficient, the dynamic
range, f, and can be offset from zero by the addition of a
constant, Eqs.2.86 and 2.87.


f1 (t ) = f1e1t + C1

f1 (t ) = f1e 1 + C1

(2.86)

f1 = 4 0 = 4

The response decays 63.2% of the dynamic range in the


first time constant, leaving 36.8% of the dynamic range
remaining.
0.632 f1 = ( 0.632)( 4) = 2.53
and
0.368 f1 = ( 0.368)( 4) = 1.47

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

86
Fig. 2.43 a Step response data
for an exponential decay. b The
time of the datum equal to 63.2%
of the initial value and 36.8% of
the remaining change to the final
value equals the time constant ,
1=0.6 sec

f1(t) 2

f1(t) 2

t, sec

f1 (t ) = f1e + C1

15

f2(t) 10

f2(t) 10

5
2

t, sec

10

0.632 f1(0+)

0.368 f1(0+)

t, sec

20

15

0
0

2
f 2 ( t ) = f 2 1 e + C2

0.368 f2(ss)

0.632 f2(ss)

t, sec

10

f 2 (t ) = 20 1 e1.5

f1 (t ) = 4 e 0.6

Example TwoDetermine the response function which


describes the data plotted in Fig.2.44.
The response is a stable exponential growth of the form
of Eq.2.87 with C2=0 since the response grows from zero.
The dynamic range is the steady-state value minus the initial
value.

f 2 = f 2 ( ss ) f 2 ( 0+ )

The time at which the response equals 1.47 is the time constant, 1=0.6sec. The response function is
t

20

0
0

Fig. 2.44 a Step response data


for a stable exponential growth,
b The time of the datum equal to
63.2% of the dynamic range plus
the initial value, and the steadystate value minus 36.8% of the
dynamic range equals the time
constant , 2=1.5 sec

f 2 = 20 0 = 20

The response grows 63.2% of the dynamic range in the


first time constant, leaving 36.8% of the dynamic range
remaining.
0.632 f 2 = ( 0.632)( 20) = 12.0
and
0.368 f 2 = ( 0.368)( 20) = 8.0
The time at which the response equals 12.0 is the time constant, 2=1.5sec. The response function is

2.9.2Non-Oscillatory Second-Order Step


Responses
A non-oscillatory second-order step response, Eq.2.88 and
Fig.2.37, is also known as an overdamped second-order
step response. An overdamped second-order step response
occurs when the two eigenvalues are both purely real, s1 = 1
and s2 = 2 . This response is difficult to characterize, because its features are less distinctive than an oscillatory step
response.


f 3 (t ) =

M3
1
1e 2t 2 e1t
1 +
1 2 2 1

(2.88)

The eigenvalues or the corresponding time constants of the


system are established by curve fitting the step response
function f3(t), Eq.2.88 to the data. If an iterative minimization routine is available, then any reasonable two initial
guesses for eigenvalues can be used. If the iteration must
be done manually, then it is worth the time to roughly estimate initial values of the eigenvalues. The larger of the two

2.9 Eigenvalues and Response Characterization


Fig. 2.45a Overdamped
second-order step response.
b Quasi-linear portion of the
initial portion of response
extended to intersect the time
axis. The time interval between
that intersection and the time
corresponding to 63.2% of
the steady-state value is the
initial estimate of the larger time
constant

Fig. 2.46a First iteration fitting


curve to overdamped second-order step response data. b Second
iteration reducing the value of
eigenvalue 2 to improve the
fit in the initial portion of the
curve. The values used to create
the data were 1 = 0.667,
2 = 1.667 , and M 3 = 30

87

30
20

20

f3(t)

f3(t)
10
0

10

2.5

t, sec

7.5

10

30

0.632 f3(ss)
0

t=

2.5

10

7.5

t, sec

30

f3(t) 20

f3(t) 20
fest (t)

fest (t)

1 = -0.549
2 = -5.49

10
0

M 3 = 81.4

2.5

time constants (the smaller eigenvalue) is estimated using


the method described above, as if the response were a firstorder step response. The quasi-linear portion of the response
curve is extended to intersect with the time axis, eliminating
the initial curve due to the second exponential term. The time
constant of second real exponential will be smaller by some
unknown magnitude. However, the lower limit of its size is
approximately one tenth of the larger time constant, if it contributes 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
the overdamped step response.
Example Determine the response function which describes
the data plotted in Fig.2.45.
The larger time constant is estimated by extending the
quasi-linear portion of the response to intersect the time
axis, Fig.2.45b. The time at which the response has grown to
63.2% of the steady-state value is then determined. The difference between these times is used as the initial estimate of
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.

1 1.82 sec and 2

30

f3(ss)=27

1.82 sec
= 0.182 sec
10

7.5

t, sec

10

1 = -0.549
2 = -2.50

10

M 3 = 37.1

2.5

t, sec

10

7.5

The overdamped step response formula, Eq.2.88, is written


in terms of eigenvalues.

1 =

1
1
1
0.549 and 2 =
=
= 5.49
1.82
2 0.182

The steady-state value of the overdamped unit step func


tion is
f 3 ( ss ) =

M3
1 +

1 2

2 1

f 3 ( ss ) =

( e
1

2 e
1

M3

1 2

The steady-state value of the step response data is 27. The


constant M3 can be estimated, using the observed steadystate value and the estimates of the eigenvalues.
M 3 = 1 2 f 3 ( ss ) M 3 = ( 0.549)( 5.49)( 27 ) = 81.4
The result, Fig.2.46a, has its greatest error early in the response, where the larger magnitude eigenvalue 2 has its effect. Reducing the magnitude to 2 = 2.50 improved the
curve fit, Fig.2.46b. A few more iterations would be sufficient for closely matching the data.

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

88

If the eigenvalues are equal, in other words, the characteristic equation has repeated roots, then the system is known as
critically damped. If the amount of damping is decreased
or energy storage in the system is increased, then the step
response will be underdamped and overshoot its final value.
The term, critical, is a misnomer, since the situation is not
critical, in the sense of producing an unstable response. It is,
in fact, difficult to discern from a step response, whether a
system slightly underdamped, critically damped, or slightly
overdamped.

1.75

1.00

fstep(t)

fimpulse(t)
0.00

-0.75
0

2.9.3 Oscillatory Second-Order Step Responses


An oscillatory second-order step response, Figs.2.31, 2.32,
2.35, and 2.40, occurs, when the eigenvalues are a complex
conjugate pair, s1 , s2 = jw , and reflects the physical situation of internal energy transfer between two independent
storagemodes persisting long enough for the step response
to overshoot its final value at least once. An oscillatory
second-order step response may have a steady-state or final
value which is either non-zero or zero. The latter is known as
an impulse response and occurs when the only term input is
differentiated. The two general forms are:



f step (t ) = Me t sin (w t + ) + C

(2.89)

f impulse (t ) = Me t sin (w t + )

(2.90)

The real component of the eigenvalue is the coefficient


in the exponent of the real exponential decay. The imaginary component of the eigenvalue is the angular frequency
of the oscillation in radians per second with the symbol, .
An oscillatory homogeneous response has two characteristic
times, the period of the oscillation T and the time constant of
1
.
the decay envelope, =

The period T of an oscillation is the duration of a complete cycle and is inversely proportional to the angular frequency,
Eq.2.88.

T period =

2 rad
rad
w
sec

(2.91)

Equation2.91 is easy to remember if one equates a cycle


with a circle. There are 2 rad in a circle and a cycle.
Measurement of the period T is illustrated in Fig.2.47.
The period T should be measured as the time between successive crossings of the steady-state value, as shown for the
impulse response. Unfortunately, crossing of the steady-state
value can be difficult to determine on an oscilloscope. The
less error-prone method is to measure the time from peak to

T
Period

T
Period
T

2T

3T

time in periods T

4T

5T

Fig. 2.47 Oscillatory second-order step and impulse responses. The


true period is measured from crossing to crossing of the steady-state
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
prone due to the ease in identifying peaks

peak of the response. The peak-to-peak time is not equal to


the period due to the decay of the oscillation, but the inherent error is less than the typical error finding the true period
using the steady-state crossing.
Instruments typically report frequency in cycles per second or hertz, which is given the symbol f. A common and
substantial error is to neglect to convert from a frequency
in hertz to the angular frequency in radians per second before performing system dynamics calculations. All system
dynamics calculations must be performed using angular frequency in radians per second.
The real component of the eigenvalue is determined from
two values of the response. In principle, any two values can
be used, other than the crossing of the steady-state value,
where the sinusoid equals zero. In practice, values at two
peaks or two troughs are used to eliminate the sinusoid
from the computation, since sin(w t + ) 1 at peaks, and
sin(w t + ) 1 at troughs. The sinusoids are only approximately equal to one and negative one due to the slight shift
of the peaks and troughs created by the decay envelope but
that error is negligible.
The differences between peak values and steady-state
are identified as a and b, between trough values and steadystate as c and d in Fig.2.48. The time between the peaks or
troughs is approximately equal to the period T. In general,
the coefficient Mis unknown until the real coefficient of the
eigenvalue is established. Using the value a at time t1 and b
at time t1+T with sin(w t + ) 1
a = Me t1 and b = Me (t1 +T )
Hence,
Me t1 Me (t1 +T ) = a b M (e t1 e (t1 +T ) ) = a b

2.9 Eigenvalues and Response Characterization

89

1.75

a
1.00

fimpulse(t)

t2

0.00

-0.75

t1

fstep(t)

should be calculated as the steady-state value minus the


trough value so that difference is positive.
Applying Eq.2.92 to step response in Fig.2.48 where
a=0.62, b=0.24, and T = 1.

c
0

0.62 0.24
ln 1

0.62
=
= 0.95
1

2.9.4 Time Step for Response Calculations


T

2T

3T

time in periods T

4T

5T

Fig. 2.48 Oscillatory second-order unit step and impulse responses.


The difference between peak and steady-state values, a and b, or the
difference between steady-state and trough values, c and d, are used to
calculate , the real component of the complex conjugate eigenvalues

e t1 e (t1 +T ) =

a b
a b
e (t1 +T ) = e t1
M
M

Divide both sides by e t1 and use the property of exponenea


tials b = e a b
e
t +T
e (1 )
t1

e t1
t1

a b
t1

Me

e T = 1

a b
t + T t
e ( 1 ) 1 = 1
Me t1
a b
Me t1

Substitute using a = Me t1
e T = 1

a b
a

Take the natural logarithm of both sides.


a b
a b
ln(e T ) = ln 1
T = ln 1

a
a
This will yield


a b
ln 1

a
=
T

(2.92)

If the peak or trough values are more than one cycle apart,
multiply the period T in the denominator of Eq.2.92 by the
number of cycles separating the values. Differences between
steady-state and trough values, such as c and d in Fig.2.48,

The real components of the eigenvalues of a system determine


the rates of exponential decays in the homogeneous response.
If a system has multiple eigenvalues, then the smallest magnitude real component min corresponds to the largest time
constant max, due to their inverse relationship, and establishes
the duration of the transient period of the response. The magnitude or absolute value of the real component is used because
is negative. A response plot should extend into the beginning
of the steady-state response of the system. A first estimate for
the duration of a plot is tmax 7 .
A time step, t or T, is the time between consecutive
calculations of a variables response, its time history. The
use of t for the difference between the times of consecutive calculations is self-explanatory. The symbol T is also
used, both for the practical reason that MATLAB and other
programming languages are restricted to the Roman characters of a computer keyboard and because it is the symbol for
the period of a periodic function. The calculation is periodic
since it is repeated at a constant time interval.
When we repeatedly evaluate a response function for successive values of time to create a plot, we want the time step
to be small enough to create a faithful representation of the
function. If the time steps are too large, then a smooth function becomes broken into a piece-wise continuous function
comprising line segments. If the time steps are too small,
then we create many more data than we need and take excessive time doing so. Using 1/200 of the smallest characteristic
time of the system, either a time constant or a period T, will
yield a good plot.
We shall see in Chap.8 that the magnitude of the time
step is critical when we use the Euler and RungeKutta algorithms to solve a set of first-order differential equations.
These algorithms are recursive, meaning that the next
value is calculated using previous values. Specifically, the
Euler and RungeKutta algorithms estimate the first derivative of output variable, multiply it by the time step to calculate the change in the output variable, and then add that
change to the present value to estimate the next value of
theoutput variable. The estimate of the first derivative of
the output variable improves as the time step is decreased.
If the time step between calculations is too large, then the

90

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

estimate of the derivative can be far enough in error to create


a numerical instability, so called because numerical instabilities yield the same shaped plots as physical instabilities.
They blow up. When we use numerical methods to solve
for the response of a dynamic system, the time step t will
be a fraction of the smallest characteristic time of the system.

2.10Laplace Transformation and Transfer


Functions
2.10.1 Laplace Transformation
A transformation is an operation which changes the independent variable of a function. The Laplace transformation
changes a function of time to a function of the Laplace variable, s. The independent variable defines the domain of a
function. Consequently, we speak of the Laplace transform
as transforming a function from the time-domain to the
Laplace-domain, in other words, to the Laplace variable,
s. The importance of the Laplace transformation is that timedomain differential equations are transformed to Laplacedomain algebraic equations.
The Laplace transformation was introduced in the context
of forming the characteristic function of a differential equation in Example Two of Sect.2.5. We introduced the Laplace
transformation by stating without proof, that it is a linear operator and, as such, the Laplace transform can be distributed
onto each term of a sum, and constants can be factored out of
the transform. We also stated that the Laplace transformation
can be viewed as an operator which transforms the differentiation with respect to time operator, d/dt, to the Laplace
variable, s, when the initial condition terms, which apply to
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
function of the Laplace variable, s, which is X(s).
Our application of the Laplace transformation will be
to transform an input function from the time-domain to the
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 transformed differential system equation into a ratio known as a
transfer function, which has the unique property of representing a systems dynamic relationship in the form of a multiplicative operator. Multiplying a systems transfer function
by the Laplace transform of an input function yields the output variables response function in the Laplace-domain. Performing the inverse Laplace transformation yields the output
variables time-domain response. We will discuss transfer
functions in the next section, but we must first develop the
Laplace transformation of functions of time.

The Laplace transformation is defined as the integral:

L { f (t )} = F ( s ) = f (t ) e

st

dt

(2.93)

The function of time to be transformed, f (t ), is multiplied


by the complex exponential, e st , and the interval is evaluated from t = to t = + . All functions f (t ) of practical
interest satisfy the necessary condition for transformability.
We have omitted the dummy variable for time. The use of
as the time constant of first-order systems can lead to confusion, when is also used as a dummy variable. It is convention to use lower case for time-domain variables and capitalize Laplace-domain variables, although it is frequently not
possible. A case in point is our use of capital F for force in
the time domain. The variable transformed to the Laplacedomain is F(s), where the functional notation (s), which is
often omitted in the time-domain, i.e., F(t), is essential to
indicate that the variable is in the Laplace-domain.
We customarily choose to define time, t=0, when we
apply an input to a system. The input function is zero for
time, t<0. Consequently, there is no need to extend the lower
limit of integration to t = . It is more appropriate to use a
one-sided Laplace transformation where the lower limit of
integration is t = 0:

L { f (t )} = F ( s ) = f (t ) e

st

dt

(2.94)

We will establish that the Laplace transformation is a linear


operator by confirming it has the two essential properties of
linear operators. First, doubling the input f(t) to the Laplace
transformation doubles the output F(s):

{2 f (t )} = 2 f (t ) e st dt = 2 f (t ) e st dt = 2 F ( s )

Second, the Laplace transformation can be distributed onto a


sum of transformable functions:

L { f (t ) + g (t )} = ( f (t ) + g (t )) e

st

dt

L { f (t ) + g (t )} = f (t ) e
0

st

dt + g (t ) e st dt
0

L { f (t ) + g (t )} = F ( s ) + G ( s )
Therefore, the Laplace transformation is a linear operator.
We will evaluate three Laplace transform integrals to reveal a few of the properties of the Laplace transform, and
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

the Laplace transform integral in practice. We will use a table


Laplace transform pairs instead, Table2.3.

91

We seek the integral,


our result by 1s :

2.10.1.1Example One: Laplace Transformation


of the Heaviside Unit Step Function
f ( t ) = us ( t )

L {u (t )} = U ( s ) = u (t ) e
s

st

dt

Recall the definition of the Heaviside unit step function:

1
1
( s ) e st dt = e st
s
s

L {u (t )} = u (t ) e
s

us (t ) = 1 for t 0
The unit step function is constant and equal to unity over the
entire integral. Consequently, we can evaluate the unit step
function first and then integrate:

u (t ) e
s

st

du deu
=
dx
dx

e st
s

du
deu
dx =
dx = eu
dx
dx

du
dx = eu
dx

st

st

d ( st )
dt

dt = e st

sdt tds
st
+

dt = e
dt
dt

st

dt
ds
st
st
s dt t dt dt = se dt = e
1
0
s e dt = e
st

st

dt = e st dt
0

{us (t )} =

L {u (t )} = se( )( )
s

e s e s0

s
s

1

s

1
us (t ) =
(2.95)
s

L{

An alternative presentation of this result is as a Laplace


transform pair:
f (t ) = us (t )

1
F ( s ) = s

2.10.1.2Example Two: Laplace Transformation


of an Exponential Decay f ( t ) = e st
The Laplace transformation integral of an exponential
decay is:

L {e } = e
at

Inspecting the left side, dt equals one. The complex varidt


able, s, is not a function of time. Hence, ds equals zero.
dt

1
dt = e st
s

1
and e0 = 1,
ea
yields the Laplace transformation of the Heaviside unit step
function.

In our case, t=x and st=u, where s is an unknown complex variable.

st

st

Using the properties of exponentials, e a =

st

Integration and differentiation are inverse operations. If one


remembers how to differentiate a function, the formula for
integration can be constructed by working backward. For example, recall how to differentiate an exponential:
eu

{us (t )} =

dt = 1e dt = e dt
st

us (t ta ) = 1 for t ta
In the present case, ta = 0. Hence,

We have what we need.


Returning to the Laplace transformation of the Heaviside
unit step function, we evaluate the integral:

us (t ta ) = 0 for t < ta

dt. Multiplying both sides of

st

at st

e dt

Begin the evaluation of the integral by combining the product of exponentials using the property, e a eb = e a + b .

L{ }

e at = e at e st dt
0

L{ }

e at = e (a + s )t dt
0

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

92

We now make use of the integration formula we derived for


the integration of the Heaviside step function,

st

dt =

e st
s

replacing the variable, s, with the sum, a+s.

( a + s )t

L {e } = e

( a + s )t

e ( a + s )t
dt =
a+s

( a + s )

e
L {e } = a + s
at

L {e }
at

e (a + s )
=
a + s

at

Rewrite the denominator in standard polynomial notation


with the constant a last to yield the result in this form:
(2.96)

The inverse Laplace transformation is the process of deriving


the time-domain function, which corresponds to a Laplacedomain function. The analytical method involves integration
in the complex plane, which is graduate-level mathematics
for most mechanical engineers. Fortunately, there is no need
to evaluate a complex integral, if the Laplace transform of
f (t ) has been evaluated yielding the Laplace transform pair,
which includes the relevant F ( s ). If we have established the
transformation from the time-domain to the Laplace-domain,
we know the corresponding reverse or inverse transformation.

L { f (t )} = F ( s )

or, as a Laplace transform pair,


f (t ) = e at

1
F (s) =
s+a

2.10.1.3Laplace Transformation of Differentiation


with Respect to Time
The Laplace transformation of a function differentiated with
respect to time is,


df (t )
+
= sF ( s ) f 0
dt

( )

(2.98)

2.10.2 The Inverse Laplace Transformation

e (a + s )0 1

a + s

at

( )

t =0

e (a + s )0
a + s

1
L {e } = s + a

( )

d n f (t )
(2.99)
= s n F (s)
dt n

t =

1
L {e } = a + s

df 0+
d 2 f (t )
2
+
s
F
s
sf
0
=

(
)

2
dt
dt

We will neglect the initial conditions terms, when performing


the Laplace transformation on a differential system equation
to form a transfer function, Sect.2.10.4. In those cases, the Laplace transformation of a function differentiated with respect
to time is

e ( a + s )t
dt =
a+s

We can now evaluate the integral:


at

The Laplace transformation of a functions second derivative with respect to time is here:

(2.97)

+
where f (0 ) is the time domain value of the function at time,
t = 0+ , the instant after a singularity input, such as a Heaviside unit step function, was applied to the system.

L {L { f (t )}} = L {F ( s )}

f (t ) =

L {F ( s )}
1

(2.100)

We will refer to a time-domain function and its corresponding Laplace-domain function as a Laplace transform pair.
We will perform Laplace transformations and inverse Laplace transformations using Laplace transform pairs. Consequently, most of the effort involved in finding an inverse
Laplace transform of an output function will not be performing the inverse transformation. It will be the algebra needed
to create an exact match of the Laplace-domain function,
F ( s ), with an existing Laplace transform pair. A frequently
needed operation is partial fraction expansion, introduced in
Sect.2.10.5.

2.10 Laplace Transformation and Transfer Functions

93

2.10.3 Final Value and Initial Value Theorems


Two useful Laplace transform theorems permit determination of time-domain values without the need to perform the
inverse Laplace transformation and evaluation of the timedomain function.
The final value theorem is


lim{ f (t )} = f ( ) = lim{ sF ( s )}
t

(2.101)

s0

The final value theorem must only be applied to functions


which have a single final value. This excludes all periodic
functions and ramp functions. Beware that the final value
theorem can be evaluated for functions which do not have a
final value.
The initial value theorem is

( )

(2.102)
lim+ { f (t )} = f 0+ = lim{ sF ( s )}
t 0

2.10.4 Transfer Functions


A differential system equation is an inputoutput relationship. It is a dynamic operator which operates on the input
to yield the output, where output is the response function of
the dynamic system. We can create a second and very useful
form of the differential system equation by performing the
Laplace transformation on it. The resulting algebraic equation is then manipulated into a transfer function, as follows.
1. Use the Laplace transform of a differential system equation, neglecting the initial condition terms, to create an
algebraic equation
b0

d 2u

du
+ b 2u =
b0 2 + b1
dt
dt

d2y

dy
+ a 2 y
a 0 2 + a 1
dt
dt

2. Factor out the transformed input and output variables


0

+ b1 s + b2 U ( s ) = a 0 s 2 + a1 s + a 2 Y ( s )

3. Create a ratio of the output variable over the input variable




A transfer function is an unusual function, because it is used in


two very different ways. A transfer function can be evaluated
as a conventional function, by assigning a value to its argument, the complex variable, s = + jw . Transfer functions
are complex functions, meaning, in the general case, the argument, s, and the result of evaluating the function are complex
numbers. The second manner of using a transfer function is as
a multiplicative dynamic operator. Multiplying a transfer function by the Laplace transform of an input function yields the
Laplace transform of the output (response) function, Fig.2.49.
We are familiar with linear inputoutput relationships in
the time domain, such as the relationship between the displacement and force of a spring, F (t ) = Kx(t ). If we rearrange this as a ratio of the output F(t) over the input x(t), we
have the form of a transfer function, where the left side is the
ratio of output over input. What remains on the right side is
a multiplicative operator.
F (t ) = Kx (t )

b0 s 2U ( s ) + b1 sU ( s ) + b2U ( s ) = a 0 s 2Y ( s ) + a1 sY ( s ) + a 2Y ( s )

a 0 s 2 + a1 s + a 2 = 0

d u
du
d y
dy
+ b1
+ b 2 u = a 0 2 + a1
+ a2 y
dt
dt
dt 2
dt

(b s

Equation2.103 is a transfer function. A transfer function


is a linear operator in the Laplace-domain. The numerator
polynomial contains the operations performed on the input,
transformed into the Laplace-domain, where the Laplace
variable, s, assumes the role of differentiation with respect
to time in the time domain. Correspondingly, the denominator 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 characteristic function of the differential equation. Setting the
characteristic function equal to zero forms the characteristic
equation of the differential equation:

Output Y ( s ) b2 s 2 + b1 s + b0
=
=
G (s)
Input
U ( s ) a2 s 2 + a1 s + a0

(2.103)

F (t )
=K
x (t )

The spring equation, in the form of a transfer function, illustrates the use of a multiplicative operator, as follows. If
the input deformation, x(t), is known and the relationship between the spring deformation and spring force is expressed
in the form of a transfer function:
x(t ) = 3 sin(2t )

and

F (t )
=K
x(t )

Then,
x (t )

F (t )
x (t )

= K 3sin ( 2t )

If we perform the Laplace transformation on the timedomain spring equation, where the unknown variables in
the time domain, x(t) and F(t), are transformed into the

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

94

U(s)

Y(s)

G(s)

Input Signal

Output Signal

Linear Operator

Fig. 2.49 A Laplace-domain block diagram, introduced in Chap. 9.


The lines represent Laplace-domain variables, (signals) U(s) and Y(s),
and the block contains the operator, the transfer function G(s)

Next, distribute the variables.


a 0 s 2Y ( s ) + a1 sY ( s ) + a 2Y ( s ) = b0 s 2U ( s ) + b1 sU ( s ) + b2U ( s )
Now, apply the inverse Laplace transformation operator to
both sides.

L {a s Y ( s ) + a sY ( s ) + a Y ( s )}
= L {b s U ( s ) + b sU ( s ) + b U ( s )}
1

unknown variables, X(s) and F(s), in the Laplace-domain,


and then create the ratio of the output function over the input
function, we have a transfer function:

L {F (t )} = L {Kx (t )}

L {F (t )} = KL { x (t )}

The inverse transform of the product of the Laplace variable,


s, and a Laplace-domain variable is the time-domain variable
differentiated with respect to time.
d y (t )
L {sY ( s )} = dt
1

Likewise, the inverse Laplace transform of the product of


the nth power of s and a Laplace-domain variable is the nth
derivative.
d n y (t )

a0

d 2 y (t )
dt 2

dy (t )
d 2 u (t )
du (t )
+ a 2 y (t ) = b0
+ b1
+ b 2 u (t )
2
dt
dt
dt

+ a1

2.10.4.2Use of Transfer Functions to Determine


Response Functions
We return to Example One of Sect.2.8.1, to illustrate the
use of transfer functions to determine the response of a system to an input, i.e., to solve a differential equation. The first
example we solved using the method of undetermined coefficients was:
6

dv
+ 3v = F (t ) where F (t ) = 8us (t )
dt

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


We form the transfer function from the differential equation by performing the Laplace transformation, and then creating the ratio to the output variable over the input variable.
Recall that the initial conditions terms are neglected when
the Laplace transformation is applied to a differential system
equation to create a transfer function.
dv

dv

L 6 dt + L {3v} = L {F (t )}
dv

L dt + 3L {v} = L {F (t )}

+ a 1 s + a 2 Y ( s ) = b 0 s + b1 s + b 2 U ( s )
2

L 6 dt + 3v = L {F (t )}

2
Y ( s ) b 0 s + b1 s + b 2
=
2
U ( s ) a 0 s + a1s + a 2

(a s

dt n

To determine the differential system equation which corresponds to a transfer function first, cross-multiply by the
denominators to eliminate the ratios.

The result is the time-domain differential equation.

L {Y ( s )} = y (t )

L {s Y ( s )} + a L {sY ( s )} + a L {Y ( s )}
= b L {s U ( s )} + b L {sU ( s )} + b L {U ( s )}

a0

2.10.4.1Inverse Laplace Transformation


of a Transfer Function
The inverse Laplace transformation of a transfer function
is simply the reverse of the process of creating a transfer
function from a differential equation. The inverse Laplace
transformation of an unknown Laplace-domain variable is
the corresponding unknown time-domain variable.

L {s Y ( s )} =

Distribute the operator, factor the coefficients out, and perform the inverse transformation.

F (s)
F ( s ) = KX ( s )
=K
X (s)

6sV ( s ) + 3V ( s ) = F ( s )
(6 s + 3)V ( s ) = F ( s )

1
V (s)
=
F ( s) 6s + 3

2.10 Laplace Transformation and Transfer Functions

95

Transfer functions are multiplicative operators. Multiplying a


transfer function by the Laplace transformation of the input
function yields the output function in the Laplace-domain. We
will use the table of Laplace transform pairs, Table2.3, to find
the Laplace transform of the input function, F (t ) = 8us (t ) .
The relevant transform pair is
f(t)

F(s)

us (t ) = 0

for

t<0

us (t ) = 1

for

t>0

V (s) =
8

L {u (t )} = s

F (s) = 8

We now formulate the Laplace-domain response function,


V(s), as the product of the Laplace transformation of the
input function, F(s), and the transfer function, V(s)/F(s).
Input ( s )
F (s)

To perform the inverse Laplace transformation, we peruse


Table2.3 for a Laplace-domain function of a similar form.
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
table entry F(s).
8
V (s) =
s ( 6 s + 3)
The numerator of V(s) is a constant. (Constants are also referred to as zero-order polynomials, since any base, including a complex base, raised to the zero power equals one,
8 = 8s 0 .) The denominator of V(s) is the product of the Laplace variable, s, and a first-order polynomial in s. The relevant transform pair is
f (t)

F(s)

1
(1 e at )
a

1
s( s + a)

8
6

V (s) =

3
6
s s +
6
6

1.33
s ( s + 0.5)

The coefficient 1.33 in the numerator of the response function can be factored out of the inverse Laplace transformation, since it is a linear operator.

L {V ( s )} = L
1

Output ( s )
= Output ( s )
Input ( s )

V (s)
8 1
= V (s)
= V (s)
s 6 s + 3
F (s)

1
8

V (s) = 6
1 s ( 6 s + 3)

6

8
V (s) =
s ( 6 s + 3)

1
s

The Laplace transformation of the input function is

L {F (t )} = L {8u (t )}

An exact match is needed to use a transform pair. Note there


is no coefficient multiplying s in the first-order factor in the
denominator of the transform pair. The coefficient must be
removed from the response function.

v (t ) = 1.33

v (t ) =

1.33
1 e 0.5t
0.5

1.33

s ( s + 0.5)

s ( s + 0.5)

v (t ) = 2.66 1 e 0.5t

Note the above yields the same result we obtained from the
method of undetermined coefficients in Sect.2.81.

2.10.5 Partial Fraction Expansion


The previous example used a first-order transfer function and
a step input. The simplicity of the transfer function and the
input required only simple algebra to create the exact match,
with a Laplace transform pair needed for the inverse Laplace
transformation. Simple algebra may not be sufficient to yield
a transform pair match, when the transfer function and/or the
input are of higher order. Although Laplace transform tables
with hundreds of transform pairs exist, Table2.3 is deliberately limited for didactic purpose, to the step response of
second-order systems. Any response function with denominator more complicated than s (a 0 s 2 + a1 s + a 2 ) will not be
found in Table2.3.
Partial fraction expansion is used to split a Laplace-domain
function (or signal) with a high-order denominator polynomial
into a sum of terms with lower-order denominators in order to
match entries in a Laplace transform table, and, thereby, determine the inverse Laplace transform. The denominators of

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

96

Expand each product on the right side.

the expansion terms are chosen to suit the engineer. The corresponding numerators are unknown, other than being polynomials of s one order lower than the denominator, and must be
found. For example,
F1 ( s ) =

N1 ( s )

s ( s + a ) s + 2w n s + w
2

2
n

xs 2 + ys + z = As 3 + A2w n s 2 + Aw n2 s

(
+ ( Bs
+ (Cs

+ Aas + Aa 2w n s + Aaw n2

A
B
Cs + D
= +
+ 2
s s + a s + 2w n s + w n2

N2 (s)

s ( s + a )( s + b )
2

s ( s + a ) s 2 + 2w n s + w n2
=

+ Cas + Ds + Das

+ Aw n2 + Aa 2w n + Bw n2 + Da s + Aaw n2
Equate the coefficients of like powers of s to create a set
of simultaneous algebraic equations. Note that there is no
cubic term on the left side, which is equivalent to a coefficient equal to zero.
s3
s2
s1
s0

A B
C
D
+ +
+
2
s s+a s+b
s

+ ys + z s ( s + a ) s 2 + 2w n s + w n2

+ ( A2w n + Aa + B 2w n + Ca + D ) s 2

0 = A+ B +C
x = A2w n + Aa + B 2w n + Ca + D
y = Aw n2 + Aa 2w n + Bw n2 + Da
z = Aaw n2

The unknowns are A, B, C, and D. Collect like unknowns and


move them to the right side of the coefficient terms.
s3
s2

0 = A+ B +C
x = ( 2w n + a ) A + 2w n B + aC + D

s1

y = w n2 + a 2w n A + w n2 B + aD

2.10.5.1 Example One: Partial Fraction Expansion


We will illustrate the technique using F1(s) above with
N1 ( s ) = xs 2 + ys + z . First, multiply both sides by the denominator of the original signal. The denominators of the
expansion terms are each a factor of the denominator of the
original signal. Consequently, the multiplication will eliminate the denominators on both sides of the equation when the
common factors are canceled.
2

+ B 2w n s 2 + Bw n2 s

xs 2 + ys + z = ( A + B + C ) s 3

The constants of the numerator terms are determined by


multiplying both sides by the denominator of the original
transfer function, canceling common factors, expanding into
a polynomial in s, and equating coefficients of like terms to
form a set of algebraic equations.

( xs

Collect like powers of s.

where, N(s), a polynomial in s, is the numerator of the original transfer function.


Note, that the second-order denominator factor is written
in terms of a damping ratio , and an ideal, undamped, natural
frequency n. This factor corresponds to a complex conjugate
pair of eigenvalues. Its inverse Laplace transform is a damped
oscillation.
There is only one important special case, when the Laplace variable, s, occurs in the denominator of the original
transfer function, as a factor raised to a power. In that case,
that denominator and all lower powers are included as terms
in the partial fraction expansion. For example, s2 is a factor
in the denominator of F2(s):
F2 ( s )

z = aw A
2
n

The set of simultaneous equations can be solved by direct


substitution and elimination, Gaussian elimination, or using
linear algebra. We will illustrate the use of linear algebra,
the fundamentals of which are reviewed in Chap.7. The left
of the set of equations is expressed as a column vector. The
right side is the product of a four by four matrix and the column vector of the unknowns.

A s ( s + a ) s 2 + 2w n s + w n2
s

) + Bs ( s + a ) ( s

+ 2w n s + w n2

s+a

) + (Cs + D ) s ( s + a ) ( s

+ 2w n s + w n2

s + 2w n s + w
2

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

2
n

2.10 Laplace Transformation and Transfer Functions

s3
2

s
s1
s0

1
0
x 2w + a
n
=
y w n2 + a 2w n

aw n2
z

0 A
1 B

a C

0 D

1
a
0
0

2w n

w n2
0

97

Division of a matrix by a matrix is not defined in linear algebra. The equivalent operation is to multiply a matrix and its
inverse to yield the identity matrix, I. The identity matrix,
I, consists of ones on the main diagonal, from top left to
lower right and zeros elsewhere. The identity matrix, I, fills
the role that the number one fills in scalar multiplication.
It can be inserted or removed as a factor without changing
a product. Matrix multiplication has restrictions. Except for
the identity matrix, I, and a matrix and its inverse, matrix
multiplication is not commutative, meaning that the position
of factors in a product cannot be changed without affecting
the result. Further, excepting only the identity matrix, I, a
factor cannot be inserted into an existing product. We must
multiply by appending a factor to the left or right end of a
product. The former is pre-multiplication and the latter is
post-multiplication.
We solve for the vector of unknowns by pre-multiplying
both sides by the inverse of the matrix on the right side.

2w + a
n

w n2 + a 2w n

aw n2

1 0
a 1
0 a

0 0

2w n

w n2
0

1
0
x 2w + a
n
=
y w n2 + a 2w n

aw n2
z

2.10.5.2Example Two: Derivation


of a Laplace Transform Pair
Our second example partial fraction expansion will also illustrate how a Laplace transform pair can be derived from
an existing transform pair. We will derive the transform pair
for the signal:
F (s) =

( s + a ) ( s + b)

using the transform pair of an exponential decay, Eq.2.95.


The first step is to express the function as the sum of two
terms:


F (s) =

( s + a )( s + b)

A
B
+
s+a s+b

(2.104)

where A and B are constants to be determined in terms of the


constants a and b. Follow the technique of multiplying both
sides by the denominator of the original signal; cancel terms;
and then expand and collect the coefficients of like powers
of s to create a set of simultaneous equations.

( s + a ) ( s + b) = A ( s + a ) ( s + b) + B ( s + a ) ( s + b)
s+a
s+b
( s + a ) ( s + b)
1
2w n

w n2
0

1 0
a 1
0 a

0 0

2w + a
n

w n2 + a 2w n

aw n2

1 0 A
a 1 B

0 a C

0 0 D

1
2w n

w n2
0

The product of a matrix and its inverse is I.


1

2w + a
n

w n2 + a 2w n

aw n2

1
2w n

w n2
0

1 0
a 1
0 a

0 0

0 1
x 0
=
y 0

z 0

0
1
0
0

1
0
1
0

0 A
0 B
0 C

1 D

which can be removed, yielding the vector of unknowns.


1

2w + a
n

w n2 + a 2w n

aw n2

1
2w n

w n2
0

1 0
a 1
0 a

0 0

0 A
x B
=
y C

z D

( s + a )( s + b)
( s + a )( s + b)

A ( s + a ) ( s + b)
s+a

B ( s + a ) ( s + b)
s+b

1 = A ( s + b) + B ( s + a )
Expand and collect the coefficients of like powers of s.
1 = As + Ab + Bs + Ba 1 = ( A + B ) s + ( Ab + Ba )
Equate coefficients of like powers of s on both sides of the
equation to yield a set of two simultaneous equations with
two unknowns, A and B.
s1 0 = A + B
s 0 1 = Ab + Ba

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

98

Solve for A and B using direct elimination, Gaussian elimination, or linear algebra. Using linear algebra, we find:
s1 0 = A + B
0 1 1 A
=

s 0 1 = bA + aB
1 b a B

f (t ) =

Combine terms and order the exponentials alphabetically by


their exponents:

Premultiply both sides by the inverse of the matrix.


1 1
b a

0 1 1
1 = b a

1 1
b a

1 1 A
b a B

f (t ) =
to yield

1 at
(e e bt )
f (t ) =
b a
Hence,

L {F ( s )} = L
1

( s + a )( s + b)

A
B
+
s+a s+b

We now use the Laplace transform pair for an exponential


decay:
1
F (s) =
s+a

L {F ( s )} = L
1

1 1

a b a b
s + a + s + b

The inverse Laplace transformation is a linear operator. It


can be distributed onto a sum.

{ F ( s )} =

a b +
s + a

a b
s + b

The numerator terms are constants and can be factored out of


the inverse Laplace transformation.
f (t ) =

1
ab

1
1

+
s + a a b

s + b

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
transformation using that transform pair.

at

e bt = f (t )

or


1 1

a b a b
+
F (s) =
s+a
s+b

f (t ) = e

+
+
s
a
s
b
)( )
(

1
L {F ( s )} = b a (e

The partial fraction expansion of F(s) is

at

1
1 1
(e bt e at ) f (t ) =
(e bt e at )

1 a b
a b

0
A
1 = I B

1
1
A a b
1 1 0 A
b a 1 = B B = 1

a b

F (s) =

1 bt
1 at
e +
e
a b
a b

1 at
1
bt
(2.105)

e e =

+
b
a
s
a
s + b)
(
)(

We have derived the Laplace transform pair:

1 at
bt
f (t ) = b a e e

1
F (s) =

( s + a )( s + b)

Again, the constants, a and b, can be purely real or complex. If they are complex conjugates, then the time-domain
expression is a decaying sinusoid. The purely real, time-domain expression is derived using the logic we used to solve
oscillatory responses with the method of undetermined coefficients. Beginning by substituting a complex conjugate pair
for a and b, a = + jw and b = jw
1 at
e e bt
f (t ) =
b a

( + jw )t
1
e ( jw )t
f (t ) =
e

+
j
j
(
)
(
)

We next use the property of exponentials, e a + b = e a eb .


f (t ) =

1 ( + jw )t
e ( jw )t
e
j 2w

f (t ) =

1 t jwt
e e
e t e jw t
j 2w

2.10 Laplace Transformation and Transfer Functions


Fig. 2.50a Spring-mass-damper
system driven by an input force,
F(t). b Plot of the input force
F (t ) = 10us (t ) + 10 1 e 0.25t

99

x,v

F(t)

25

20

F(t) 15
N 10

We now factor out the purely real exponential, and group the
purely imaginary exponentials, the complex exponential unit
vectors, to use Eulers sine formula.
f (t ) =

e jw t e jw t
1 t e jw t e jw t
e t

(
)
=
e
f
t

w
j2
w
j2

f (t ) =

e t sin(w t )

1
1 t

e sin (w t ) =
w
( s + + jw )( s + jw )

10

t, sec

15

20

Begin by creating the transfer function of the system


equation. Perform the Laplace transformation on both sides
of the system equation, neglecting the initial condition terms.

dF

(2.106)

dF
=
dt

d 2v

dv
M 2 + b + Kv
dt
dt

d 2v
+
2

L dt = L M dt

dv

L b dt + L {Kv}

Factor out the parameter terms, which are constants.

or


Distribute the Laplace transformation operator onto each


term of the sum.

Hence,


1
1 t

e sin (w t ) = 2
w
+
+ 2 + w2
s
2
s

dF

(2.107)

2.10.5.3Example Three: Response


of a Spring-Mass-Damper System
This example illustrates the use of transfer functions and partial fraction expansion, to find the response of an oscillatory
second-order system to an input which is superposition of a
step and a stable first-order exponential growth. Determine
the velocity of the mass of the spring-mass-damper system
shown in Fig.2.50a. The system equation relating the input
force to the velocity of the mass is

d 2v
+b
2

L dt = M L dt

dv

L dt + KL {v}

The Laplace transformation of an unknown time-domain


variable is an unknown Laplace-domain variable. Differentiation with respect to time transforms to the Laplace variable, s.
sF ( s ) = Ms 2V ( s ) + bsV ( s ) + kV ( s )
Collect the transformed output variable V(s). A transfer
function is the ratio of the output over the input:

sF ( s ) = Ms 2 + bs + k V ( s )

dF
d 2v
dv
= M 2 + b + Kv
dt
dt
dt

Output V ( s )
s
=
=
Input
F ( s ) Ms 2 + bs + k

=
6. The force input is
with b=1, M=2, and K
F (t ) = 10us (t ) + 10(1 e 0.25t ) , Fig.2.50b.

Note that the operators from the output variable side of the
system equation are in the dominator of the transfer function.

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

100

The standard form for a transfer function requires the coefficient of the highest-order term of the denominator to be
cleared. This form corresponds to the standard form of a differential equation.
1
s
V (s)
M
=
k
b
F (s)
s+
s2 +
M
M

Factor

L {F (t )} = 10L {u (t )} + 2.5L 0.25 (1 e


s

Substitute in the parameter values.


V (s)
=
F (s)

The relevant Laplace transform pairs from Table2.3 are:


f (t)

F(s)

us (t ) = 0

for

t<0

us (t ) = 1

for

t>0

0.25t

L {F (t )} = L {10u (t )} + L {10 (1 e
s

)}

0.25t

)}

We must create exact matches with the time-domain functions of the Laplace transform pairs. Factor the coefficient 10
out of both the transformations. Multiply the second transform by the unity ratio, 0.25/0.25.

L {F (t )} = 10L {u (t )} + 10 0.25 L {(1 e


0.25

10
2.5
+
s s ( s + 0.25)

V ( s ) 10
2.5
0.5s

=
+
= V (s)

F ( s ) s s ( s + 0.25) s 2 + 0.5s + 3

Cancel the Laplace variable, s, where it occurs in both the


numerator and denominator.


10
0.5 s
2.5
0.5 s
s s 2 + 0.5s + 3 + s s + 0.25 s 2 + 0.5s + 3 = V ( s )
)
(

Apply the Laplace transformation operator to both sides, and


distribute to the terms of the right side.
s

F (s) =

5
1.25
+
= V (s)
s + 0.5s + 3 ( s + 0.25) s 2 + 0.5s + 3

1
s(s + a)

L {F (t )} = L {10u (t ) + 10 (1 e

The Laplace transform of the output variable is

1
s

1
(1 e at )
a

0.25t

F (s)

Next, determine the Laplace transformation of the input:


F (t ) = 10us (t ) + 10 1 e

A transfer function is a multiplicative operator. Multiply the


transfer function by the Laplace transform of the input to
yield the Laplace transform of the output.

1
s
V (s)
0.5s
2

= 2
1
6
F
s
0.5s + 3
s
+
()
s2 + s +
2
2

0.25t

0.25t

We now have the exact matches needed to perform the transformations using the Laplace transform pairs.

L { f (t )} = 10L {u (t )} + 2.5L 0.25 (1 e


s

1
into the second transform:
0.25

0.25t

)}

We now wish to perform the inverse Laplace transformation to return to the time-domain. Note the denominators
of the two terms contain the common factor, second-order
factor, s 2 + 0.5s + 3 . This is the characteristic function of
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 homogeneous or natural response. If the eigenvalues of the system
are real, then the homogeneous response is not oscillatory.
If the eigenvalues are a complex conjugate pair, then the response is oscillatory.
Set the characteristic function equal to zero, to form the
characteristic equation and solve it using the quadratic equation.
s 2 + 0.5s + 3 = 0 s1 , s2 =

0.5

s1 , s2 = 0.25 j1.71

(0.5)2 ( 4)(3)
2

2.10 Laplace Transformation and Transfer Functions

101

The eigenvalues s1 , s2 = 0.25 j1.71 are a complex conjugate pair. Therefore, the homogeneous or natural response of
the system is oscillatory or underdamped. Accordingly, we
must use Laplace transform pairs which correspond to oscillatory time-domain responses. These are simple to identify
in Table2.3; the time-domain response contains a sinusoid.
The denominator of the corresponding Laplace-domain signal is written in terms of the ideal, undamped natural frequency, n, and the damping ratio, . Specifically, the two
terms of the output

nator of the second term into a summation with denominators that match Laplace transform pairs. The denominators
of the partial fraction expansion terms are factors of the left
side. The numerators of the expansion terms are polynomials
in s, one order lower than their denominators. Theconstants,
A, B, and C must be determined.
1.25

( s + 0.25) ( s

1.25 ( s + 0.25) s 2 + 0.5s + 3

( s + 0.25) ( s 2 + 0.5s + 3)

+ 0.5s + 3

) + ( Bs + C )( s + 0.25) ( s

A ( s + 0.25) s 2 + 0.5s + 3
s + 0.25

A
Bs + C
+ 2
s + 0.25 s + 0.5s + 3

Multiply both sides by the denominator of the left side,


and cancel factors which appear in the numerator and
denominator.

5
1.25
+
= V (s)
s 2 + 0.5s + 3 ( s + 0.25) s 2 + 0.5s + 3

+ 0.5s + 3

s + 0.5s + 3

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

correspond to

Distribute the unknown constants.

) (

1.25 = As 2 + A0.5s + A3 + Bs 2 + Bs 0.25 + (Cs + C 0.25)

5
5
= 2
s + 0.5s + 3 s + 2w 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

( s + 0.25) ( s

) (s + a)(s

+ 0.5s + 3

1.25
2

+ 2w n s + w n2

Checking the forms of the denominators of the terms against


the forms of the denominators of the transform table pairs,
we find a match for the first term. The relevant transform
pair is
f (t )

wn
1 2

F (s)
e

w nt

sin(w n 1 t )
2

w n2
2
s + 2w n s + w n2

There is not a transform pair table match for the second term.
We must use partial fraction expansion to break the denomi-

1
0
1
0.5 0.25
1

3
0
0.25

Create a set of three equations, by equating the coefficients


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.
s2
0 = A+ B
1
s
0 = 0.5 A + 0.25 B + C
0
s 1.25 = 3 A + 0.25C
We will solve for the constants with linear algebra. First, express the set of equations as a single vector-matrix equation.
1
0 A
0 1
0 = 0.5 0.25
B
1



1.25 3
0
0.25 C
Premultiply both sides by the inverse of the matrix.

1
0
0 1
0 = 0.5 0.25
1


1.25 3
0
0.25

1
0 A
1
0.5 0.25
1 B

3
0
0.25 C

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

102

5
0.426
0.426 s
0.106
+
2
2
= V (s)
s + 0.5s + 3 s + 0.25 s + 0.5s + 3 s + 0.5s + 3

The product of a matrix and its inverse equals the identity


matrix, I, which can be removed.
1
0
1
0.5 0.25
1

0
0.25
3

Inspecting the output signal, we notice that the first and last
terms have a common denominator and can be summed,
eliminating a term.

0 1 0 0 A



0 = 0 1 0 B
1.25 0 0 1 C

1
0
1
0.5 0.25
1

0
0.25
3
1
0
1
0.5 0.25
1

0
0.25
3

4.894
0.426
0.426 s
+
2
= V (s)
s + 0.5s + 3 s + 0.25 s + 0.5s + 3
2

0
A
0 = I B


1.25
C

0 A


0 = B
1.25 C

The output variable, V(s), is now a sum containing three different transform pairs which are in the Laplace transform table,
Table2.3. The relevant Laplace transform pairs are:

0.426 A


0.426 = B
0.106 C

f (t )

F (s)

e at

1
s+a

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


expansion:

1.25

( s + 0.25) ( s 2 + 0.5s + 3)
1.25

( s + 0.25) ( s 2 + 0.5s + 3)

A
Bs + C
+
s + 0.25 s 2 + 0.5s + 3

e nt sin n 1 2 t

1
1 2

n2
s + 2 n s + n2
2

e nt sin n 1 2 t

s
s 2 + 2 n s + n2

1 2
d
= tan 1

= tan 1

0.426
0.426 s 0.106
+
s + 0.25
s 2 + 0.5s + 3

L {V ( s )} = L

Again, we must make exact matches against the table transform pairs to perform the inverse Laplace transformation
using the transform pairs.

4.894
2
+
s + 0.5s + 3

0.426

s + 0.25

0.426 s
2

s + 0.5s + 3

Factor out the numerator coefficients.


v (t ) = 4.894

2
+ 0.426
s + 0.5s + 3

The second term has a first-order numerator. Write the second term as two terms.
1.25

( s + 0.25) ( s 2 + 0.5s + 3)
=

0.426
0.426 s
0.106
2
2
s + 0.25 s + 0.5s + 3 s + 0.5s + 3

Now, assemble the output signal, V(s).


5
1.25
+
= V (s)
s 2 + 0.5s + 3 ( s + 0.25) s 2 + 0.5s + 3

0.426
s + 0.25

s + 0.5s + 3

The second and third terms match their Laplace transform


pair. Multiply the first term by the unity ratio, 3/3.
3
v (t ) = 4.894
3

2
+ 0.426
s + 0.5s + 3
s

1
0.426
2

s
+
0.5
s
+
3

s + 0.25

Factor three into the numerator of the first term and evaluate
the remaining ratio.

2.10 Laplace Transformation and Transfer Functions

v (t ) = 1.63

103

2
+ 0.426
s + 0.5s + 3

To evaluate the oscillatory time-domain terms, we must calculate the systems ideal, undamped, natural frequency n
and the damping ratio . The simplest method is to equate
coefficients of like powers of s of the denominators of an
oscillatory term, and the corresponding Laplace-transform
table expression.
s 2 + 0.5s + 3 = s 2 + 2w n s + w n2
First, calculate the ideal, undamped, natural frequency n,
which is the square root of the constant term.

w n = 3 = 1.73
Then, using the value of the natural frequency n, calculate
the damping ratio .
0.5
=
2w n

0.5 = 2w n

v1 (t ) = 1.63

1.73
1 ( 0.144)

s + 0.5s + 3

e (0.144)(1.73)t sin 1.73 1 ( 0.144) t


2

This term happens to be the Laplace transform pair we derived as Example Two. The exponent is the real component
of the eigenvalues, s1 , s2 = 0.25 j1.71 . The discrepancy
is due to numerical precision. The frequency 1.71 is the magnitude of the imaginary component of the eigenvalues. The
units are radians per second. Again, eigenvalues are the characteristic values of a system. They do indeed characterize the
systems natural or homogeneous response, which is excited
by any change in the input to the system.
The inverse Laplace transformation of the second term is
1

s + 0.25

v2 (t ) = 0.426 e 0.25t

The Laplace variable, s, in the numerator of the third terms


creates a constant , known as the phase shift, which is subtracted from the time-varying angle of the sine, w n 1 2 t .

s
1

e w t sin w n 1 2 t
2
= 0.426
s + 0.5s + 3
1 2

1
2
e (0.144)(1.73)t sin 1.73 1 ( 0.144) t
v3 (t ) = 0.426
2

1 ( 0.144)

where the phase shift is


1 2
w
= tan d = tan 1

to or greater than one, then the factor is not oscillatory, and a


different Laplace transform pair must be used.
We now use the Laplace transform pair table to perform
the inverse Laplace transformation of each term. The first
term yields:
v1 (t ) = 1.63

v1 (t ) = 1.63

wn
1 2

s + 0.5s + 3

w n t

1 ( 0.144)2
= 1.43 rad
= tan
0.144

Yielding
v3 (t ) = 0.430 e 0.249t sin (1.71t 1.43)
The time-domain response function plotted in Fig.2.51, is
v (t ) = v1 (t ) + v2 (t ) + v3 (t )

sin w n 1 t
2

v1 (t ) = 2.85e 0.249t sin (1.71t )

The damping ratio of an oscillatory factor is in the range of


zero to one, 0 < < 1. If the calculated damping ratio is equal

0.426
s + 0.25

v2 (t ) = 0.426

= 0.144

v3 (t ) = 0.426L

v (t ) = 2.85e 0.249t sin (1.71t ) + 0.426 e 0.25t

+ 0.430 e 0.249t sin (1.71t 1.43)

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

104

3
2

m
v(t), ___
sec

-2

10

t, sec

15

20

Fig. 2.51 Response v(t) of the spring-mass-damper system to the input


force

Problems
Problem 2.1 Use the method of undetermined coefficients
to solve the ordinary differential equations with constant coefficients below. Plot the homogeneous, particular, general
solution on the same plot using Mathcad or MATLAB.

2.1.b

dF (t )
dv (t )
=3
+ v (t )
dt
dt
dv (t )
F (t ) = 3
+ v (t )
dt

2.1.c

F (t ) = 3

2.1.d

F (t ) = 3

2.3.b

2.3.c

3F ( t ) = 9

2.3.d

3F ( t ) = 9

dv (t )
+ v (t )
dt
dv (t )
+ v (t )
dt

v 0+ = 7

F (t ) = 10 us (t )

v 0+ = 0

F (t ) = 10 us (t )

v 0+ = 6

F (t ) = 6 u s (t )

v 0+ = 10

( )
( )
( )

2 F (t ) = 8

2.2.b

2 F (t ) = 8

2.2.c

2.2.d

2 F (t ) = 8

dv (t )
+ 3v (t )
dt
dv (t )
+ 3v (t )
dt

dF (t )
dv (t )
=8
+ 3v (t )
dt
dt
dv (t )
+ 3v (t )
dt

F (t ) = 12 us (t )

( )
( )

v 0+ = 0

F (t ) = 12 us (t )

v 0+ = 9

F (t ) = 6 u s (t )

v 0+ = 18

( )
( )

Problem 2.3 Use the method of undetermined coefficients


to solve the ordinary differential equations with constant coefficients below. Plot the homogeneous, particular, general
solution on the same plot using Mathcad or MATLAB.

dF (t )
dv (t )
=9
+ 6v ( t )
dt
dt
dv (t )
+ 6v ( t )
dt
dv (t )
+ 6v ( t )
dt

( )

F (t ) = 14 us (t )

v 0+ = 0

F (t ) = 14 us (t )

v 0+ = 34

F (t ) = 14 us (t )

v 0+ = 34

F (t ) = 14 us (t )

v 0+ = 22

2.4.a

dF (t ) d 2v (t )
dv (t )
=
+3
+ v (t )
dt
dt 2
dt

2.4.b

F (t ) =

d 2 v (t )
dv (t )
+3
+ v (t )
dt 2
dt

v 0+ = 0

2.4.c

F (t ) =

d 2 v (t )
dv (t )
+3
+ v (t )
dt 2
dt

v 0+ = 6

2.4.d

F (t ) =

d 2 v (t )
dv (t )
+3
+ v (t )
2
dt
dt

v 0+ = 10

( )
( )
( )

( ) = 14

( )

d v 0+

( )

d v 0+

( )

d v 0+

( )

d v 0+

v 0+ = 7

dt

( )=7

dt

( )=0

dt

( )=3

dt

Problem 2.5 Use the method of undetermined coefficients to


solve the ordinary differential equations with constant coefficients below for the input F (t ) = 12 us (t ). Plot the homogeneous, particular, general solution on the same plot using
Mathcad or MATLAB.
2.5.a

2 F (t ) =

d 2 v (t )
dv (t )
+8
+ 3v (t )
dt 2
dt

v 0+ = 5

2.5.b

2 F (t ) =

d 2 v (t )
dv (t )
+8
+ 3v (t )
dt 2
dt

v 0+ = 0

2.5.c

2.5.d

2 F (t ) =

v 0+ = 5

F (t ) = 12 us (t )

dv (t )
+ 6v ( t )
dt

Problem 2.4 Use the method of undetermined coefficients to


solve the ordinary differential equations with constant coefficients below for the input F (t ) = 10 us (t ). Plot the homogeneous, particular, general solution on the same plot using
Mathcad or MATLAB.

( )

F (t ) = 10 us (t )

Problem 2.2 Use the method of undetermined coefficients


to solve the ordinary differential equations with constant coefficients below. Plot the homogeneous, particular, general
solution on the same plot using Mathcad or MATLAB.
2.2.a

3F ( t ) = 9

-1

2.1.a

2.3.a

d v 0+

( )

d v 0+

dF (t ) d 2 v (t )
dv (t )
+
=
+8
+ 3v (t ) v 0 = 9
dt
dt 2
dt
d 2 v (t )
dv (t )
+8
+ 3v (t )
dt 2
dt

( ) = 18

( )

dt

( ) = 12

dt

( )=0

( )

d v 0+

( )

d v 0+

v 0+ = 18

dt

( )=9

dt

Problem 2.6 Use the method of undetermined coefficients to


solve the ordinary differential equations with constant coefficients below for the input F (t ) = 14 us (t ). Plot the homogeneous, particular, general solution on the same plot using
Mathcad or MATLAB.
2.6.a 3F (t ) =
2.6.b 3

d 2 v (t )
dv (t )
+9
+ 6v ( t )
dt 2
dt

( )

v 0+ = 0

dF (t ) d 2 v (t )
dv (t )
+
=
+9
+ 6v (t ) v 0 = 34
2
dt
dt
dt

2.6.c 3F (t ) =

d 2 v (t )
dv (t )
+9
+ 6v ( t )
2
dt
dt

( ) = 18

d v 0+
dt

( ) = 12

( )

d v 0+

( )

d v 0+

v 0+ = 34

dt

( )=0

dt

Chapter 2 Appendix
2.6.d 3F (t ) =

d 2 v (t )
dv (t )
+9
+ 6v ( t )
dt 2
dt

105

( )

v 0+ = 22

( )=9

d v 0+
dt

Problem 2.7 Use the method of undetermined coefficients to


solve the ordinary differential equations with constant coefficients below for the input F (t ) = 10 us (t ). Plot the homogeneous, particular, general solution on the same plot using
Mathcad or MATLAB.
2.7.a

dF (t ) d 2 v (t )
dv (t )
+
=
+3
+ 5v (t ) v 0 = 7
dt
dt 2
dt

2.7.b

F (t ) =

d 2 v (t )
dv (t )
+3
+ 5v (t )
dt 2
dt

v 0+ = 0

2.7.c

F (t ) =

d 2 v (t )
dv (t )
+3
+ 5v (t )
dt 2
dt

v 0+ = 6

F (t ) =

d v (t )
dv (t )
+3
+ 5v (t )
dt 2
dt

v 0+ = 10

2.7.d

( ) = 14

( )

d v 0+

( )

d v 0+

( )

d v 0+

( )

d v 0+

dt

2 F (t ) =

d 2 v (t )
dv (t )
+3
+ 8v (t )
dt 2
dt

v 0+ = 5

2.8.b

2 F (t ) =

d 2 v (t )
dv (t )
+3
+ 8v (t )
dt 2
dt

v 0+ = 0

2.8.c

2.8.d

2 F (t ) =

dF (t ) d 2 v (t )
dv (t )
=
+3
+ 8v (t )
2
dt
dt
dt

( )=0

dt

( )=3

dt

d 2 v (t )
dv (t )
+3
+ 8v (t )
dt 2
dt

( ) = 18

( )

d v 0+

( )

d v 0+

( )

d v 0+

( )

d v 0+

v 0+ = 9
v 0+ = 18

dt

d 2 v (t )
dv (t )
+9
+ 6v ( t )
dt 2
dt

( )

3F ( t ) =

2.9.b

2.9.c

3F ( t ) =

d 2 v (t )
dv (t )
+5
+ 9v ( t )
dt 2
dt

v 0+ = 34

2.9.d

3F ( t ) =

d 2 v (t )
dv (t )
+5
+ 9v ( t )
dt 2
dt

v 0+ = 22

v 0+ = 0

dF (t ) d 2 v (t )
dv (t )
+
=
+5
+ 9v (t ) v 0 = 34
dt
dt 2
dt

2 F (t ) =

2.10.b

F (t ) =

2.10.c

2 F (t ) =

2.10.d

3F ( t ) =

( ) = 12

d 2 v (t )
dv (t )
+8
+ 3v (t )
dt 2
dt

d 2 v (t )
dv (t )
+3
+ 5v (t )
2
dt
dt

d 2 v (t )
dv (t )
+4
+ 9v ( t )
dt 2
dt

d 2 v (t )
dv (t )
+5
+ 18v (t )
dt 2
dt

Problem 2.11 Use Laplace transforms and transfer functions


to solve the ordinary differential equations with constant coeft

ficients below for the input F (t ) = 10 us (t ) 2 + 3 1 e 2 .

Plot the solution using Mathcad or MATLAB.


2.11.a

2 F (t ) =

2.11.b

F (t ) =

2.11.c

2 F (t ) =

2.11.d

3F ( t ) =

d 2 v (t )
dv (t )
+8
+ 3v (t )
dt 2
dt

d 2 v (t )
dv (t )
+3
+ 5v (t )
2
dt
dt

d 2 v (t )
dv (t )
+3
+ 8v (t )
dt 2
dt

d 2 v (t )
dv (t )
+5
+ 9v ( t )
dt 2
dt

dt

( )=0

dt

( )=9

dt

Problem 2.9 Use the method of undetermined coefficients to


solve the ordinary differential equations with constant coefficients below for the input F (t ) = 14 us (t ). Plot the homogeneous, particular, general solution on the same plot using
Mathcad or MATLAB.
2.9.a

2.10.a

( )=7

dt

Problem 2.8 Use the method of undetermined coefficients to


solve the ordinary differential equations with constant coefficients below for the input F (t ) = 12 us (t ). Plot the homogeneous, particular, general solution on the same plot using
Mathcad or MATLAB.
2.8.a

stant coefficients below for the input F (t ) = 10 us (t ) . Plot


the solution using Mathcad or MATLAB.

( ) = 18

d v 0+
dt

( ) = 12

( )

d v 0+

( )

d v 0+

( )

d v 0+

Chapter 2 Appendix
Table 2.3 Laplace Transform Pairs
f(t)

(t ) = 0 for t 0
Unit impulse:
(Dirac delta function) (t ) = for t = 0
Unit step:
Unit ramp:

us (t ) = 0

for

t<0

us (t ) = 1

for

t>0

ur (t ) = 0

for

t<0

ur (t ) = t

for

t>0

( )=0
( )=9

dt

Problem 2.10Use Laplace transforms and transfer functions to solve the ordinary differential equations with con-

1
1 e at
a

1
s2

1
be bt ae at
ba

1
s

1
s (s + a)

1
e at e bt
ba

1
s+a

e at

dt
dt

F(s)

( s + a )( s + b)

( s + a )( s + b)

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

106
Table 2.3(continued)
f(t)

F(s)

1
1

1+
be at ae bt
ab a b

e nt sin n 1 2 t

1 2

1
1 2

1
s ( s + a )( s + b )

n2
s + 2 n s + n2
2

e nt sin n 1 2 t

1
d
1
= tan

= tan 1
1

1
1

e nt sin n 1 2 t +

1 2

= tan d = tan 1

sin ( t )
cos ( t )

s
s 2 + 2 n s + n2

purpose computer programming languages in use during that


early period were written to implement mathematical algorithms for engineers and scientists. MathWorks foresaw the
time when general purpose programming languages would
evolve into object-oriented languages to manage the execution of enormous programs, and no longer permitted
simple programming of mathematical algorithms. MATLAB
is a procedural language intended to encode algorithms.
The syntax and execution of MATLAB resemble that of the
original BASIC computer language from the 1960s, which
was developed as an introductory programming language.

Plotting in Mathcad
n2
s ( s 2 + 2 n s + n2 )

s2 + 2
s
s + 2
2

Mathcad and MATLAB


In the early 1980s, shortly after personal computers were introduced, there was a flurry of activity to develop software
for the emerging market. Two Massachusetts companies targeted engineering and scientific computation with radically
different approaches. Mathsoft, the originator of Mathcad,
chose to develop software which emulated the mixture of
mathematics, graphics, and text, which typifies manual engineering calculations. A Mathcad worksheet presents equations in as close to standard mathematical notation as possible. What the user sees is actually a graphical interface defined by user created regions or objects that support mathematics, graphics, or text. The regions within the worksheet
can be dragged around, resized, copied, and edited. Mathcad
worksheets are evaluated automatically from top to bottom,
left to right when any change is made.
MathWorks, the developer of MATLAB took a different approach from Mathsoft. Rather than creating an automated engineering computation worksheet, MathWorks
chose to create a high-level programming language with
the functions and functionality needed for engineering and
scientific computing. A high-level programming language
performs many of the tasks needed to translate an algorithm
expressed as human-readable computer code into the instructions required by the computers processor to execute
the program. Computer programs from 1950s through the
early 1980s were tiny, by todays standards. Most general

Mathcad Prime is engineering computational software designed to resemble the layout and appearance of manual
computations. Equations, graphics, and text can be placed at
will in the worksheet, with the only restriction that constants
and functions must be defined above where they are used,
because the worksheet is computed from top to bottom, left
to right. Mathcad strives to use conventional mathematical
notation, making the worksheet significantly easier to read
than computer code. Plotting a function in Mathcad requires
little more than defining the function to be plotted using an
assignment statement.

Mathcad Assignment Statements


Assignment statements are instructions to a program to assign a value or expression to a variable. Assignment statements are read from right to left. In other words, the quantity
on the right side of an assignment statement is assigned to
the variable or function on the right. Although most computer languages, including MATLAB, use the convention,
equal sign i.e., =, as the assignment operator, Mathcad
does not. The Mathcad assignment operator is :=, which is
created by typing a colon. It is supposed to connote an arrow
pointing from right to left. The reason Mathcad uses := rather than=as the assignment operator is because assignment
statements are not equalities. Assignment statements are
similar to a store command on a calculator. An assignment
statement is an instruction to store the value or expression on
the right within the variable or function name on the left. For
example, the following are three valid Mathcad assignment
statements.
a: =1
a : = a +1

F(t) : = 10 sin ( 2 t )

The first and third assignment statements also make sense


as equalities. However, the second statement written as an
equality, a = a + 1, is nonsense.

Mathcad and MATLAB

107

In Mathcad, a conventional equal sign is a command to


evaluate a variable or an expression, and display the result.
The use of an equal sign makes sense in this context, because
the quantities on either side of the equal sign are equal.

Plotting in Mathcad
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
place the plot. When you click on an empty area of a Mathcad
worksheet and begin to type, the program presumes that
you are entering an equation. If, in fact, you are typing text,
Mathcad will recognize you are entering text by the space
entered between letters. There are no spaces in equations.
When entering an equation, the space bar is used in lieu of
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.
Click in an empty area of a Mathcad worksheet and type
v(t) : 2.66*(1 e ^ 0.5* t

Space Bar Space Bar

Then either type tab, or click outside of the equation object.


You will see the following:

v(t) : = 2.66 1 e 0.5t

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


different variables. Greek characters can be entered using
the Greek alphabet pallet (produced by clicking on the
button), or by typing the combination of the Ctrl key and g
simultaneously, when the cursor is immediately to the right
of the Roman character one wishes to change to Greek.
The duration of a response plot should be six or seven
times the largest time constant in the system, Sect.2.9.4. We
are plotting a first-order response, where the time constant is
the inverse of the magnitude of coefficient in the exponent,
= 0.5. Type
t Ctrl+g : | 1/ 0.5

Simultaneously

Press the Ctrl and g keys simultaneously. Type tab or click


outside the equation object and you will see

:=

1
0.5

To insert an xy plot into a Mathcad worksheet, first place


the cursor outside an existing Mathcad text or math region.
One can then either use icons or the menus. The first time
you click on a plot icon, Mathcad displays the plotting pallet or button bar. Clicking on the icon for an xy plot in the

Upper
Limit
Dependent
Variable
Lower
Limit
Lower
Limit

Independent
Variable

Upper
Limit

Fig. A2.1 A Mathcad xy plot object showing the black rectangular


place holders. The variables can be scalars, vectors, expressions, or
functions. The limits can be numerical values or expressions

2.658
v( t)

0.689

3
2.5
2
1.5
1
0.5

0
0

10

Fig. A2.2 A Mathcad plot of the function,v (t ) = 2.66 1 e 0.5t

button bar inserts an empty xy plot into the worksheet. To


use the menu, click on the Insert menu, and follow the drop
down menu to the item XY Plot.
Insert
Graph

XY Plot.

An empty xy plot will display with solid black rectangles


termed place holders for the independent and dependent
variables and the axes limits, Fig. A2.1.
To plot v(t) vs. t, enter t in the independent variable
place holder on the x-axis, 0 in the lower limit x-axis place
holder, 7*t Ctrl+g (to see 7 ) in the upper limit x-axis
place holder, and either the function, v(t), or the expression,
2.66 (1 e 0.5t ), in the dependent place holder on the y-axis.
Clicking outside the region causes Mathcad to display the
plot, Fig. A2.2.
Mathcad autoscales axes, when the limit place holders
are blank. Often, as in this example, autoscaling leads to the
display of only a portion of the response. The limits can be
edited, after the plot is produced to change either the vertical

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

108

2.658
2.66

velocity, m/sec

2
v( t)
1
0

0
0

10
t

v( t)
0.5 t

e
2.66

2.66

2
0

velocity, m/sec

v( t)

v( t)
1

392.119
0
0

10

time, sec

Fig. A2.4 The Mathcad plot of the function, v (t ) = 2.66 1 e


matted adding gridlines, axes labels, and a plot title

0.5t

100
0
100
200
300
400
10
10

) , for-

or horizontal range shown. Clicking on the lower limit of


the vertical axis 0.689 introduces the shaped cursor into
that region. Editing 0.689 to read 0, and then either typing
tab, or clicking outside the plot region causes Mathcad to
re-evaluate and display the plot, Fig. A2.3.
Plot regions can be dragged around the worksheet, and
the frame of a plot region can be dragged to resize the plot.
Plots can also be formatted to add grid lines, change the
grid spacing, and to change the width, color, and type of line,
Fig. A2.4. Right clicking within a plot region brings up a
context-sensitive menu which includes the three items, Format, Trace, and Zoom The format dialog box has five
tabs which are self-explanatory, with the exception of the
secondary Y-axis. Clicking the check box Enable second
Y-axis produces a set of place holders on the right side of
the plot, the middle place holder is for a second dependent
variable, and the other two are the limits of the secondary Yaxis. We will find a secondary Y-axis useful, since the power
variables of energetic systems have different units and, importantly, different magnitudes. Plotting the responses of
two different power variables on one axis can lead to one
response appearing flat, since its vertical range is misscaled.
When we wish to plot two or more traces on the same
axis, typing a comma after entering a variable or expression

0.5t
Fig. A2.5 The Mathcad plot of the function, v(t ) = 2.66(1 e ), the
0.5t
expression, 2.66e , and the constant, 2.66

2.642

2.658

10
t

time, sec

Response Function v(t)

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

Response Function v(t)

0
t

10
10

Fig. A2.6 The Mathcad plot of the function,v(t ) = 2.66(1 e 0.5t ), autoscaled from t=10 to t=+10

in a place holder produces another place holder. For example, we plot v(t ), 2.66e 0.5t , and 2.66 on the same axis, by
typing a comma after entering v(t ), and again after entering
2.66e 0.5t , Fig. A2.5.
We chose to set the limits of the time axis at zero and six
or seven time constants. If we create a plot by editing the
place holders for the independent and dependent variables,
but leave the axes limits blank, Mathcad will autoscale the
horizontal axis from 10 to +10, and evaluate the dependent
variable within those limits. Since the input was applied at
time, t=0, the plot will show a response before the input
acted on the system, Fig. A2.6. Even though we did not intend for the function to be evaluated for negative time, it
can be. The polite term for the resulting plot is non-causal
since it violates cause and effect. The more common terms
include nonsense and garbage.
Do not plot negative time, unless (1) the input is applied
at the negative time of the lower limit, or (2) the response
function is multiplied by the Heaviside unit step function to
zero out the response function, until the time the input is applied. Mathcads notation for the Heaviside unit step function is capital phi, ( ). The f(x) button brings up a dialog
with all of Mathcads built-in functions. The Heaviside unit
step is in the Piecewise Continuous submenu, or can be
found in the alphabetical list. Multiplying the response func-

Mathcad and MATLAB

109

Response Function v2(t)


4.5

v, m/sec

( t) v( t) 1
0
1
10

10

Recall the exponent of the real exponential is , the real component of the eigenvalues of the system, and the frequency
is the magnitude of the imaginary component of the eigenvalues.
s1 , s2 = 0.16 j 0.68

Create an assignment statement defining the response variable, v2(t). Note the subscript 2, which is part of the functions name. Mathcad refers to a subscript which is part of
a variable of function name as a literal subscript, to distinguish it from a subscript which represents the index of a
vector. A literal subscript is created by typing a period immediately before the literal subscript. A vector subscript or
index is created by typing a left square bracket [ immediately
before the subscript. Type

v 2 ( t ) : = 2.74 e 0.16t cos ( 0.68 t 2.91) + 2.66


The time constant, which scales the duration of the plot, is
the time constant of the decay envelope. The upper limit of
the time axis should be six or seven , where
1

1
0.16

10

20

t, sec

30

6 .2

Fig. A2.8 The Mathcad plot of the response function, v2(t), formatted
with gridlines, axes labels, and a plot title

Mathcad permits assignment statements to be evaluated.


Type
t Ctrl+g .2 : | 1/ 0.16 =

Simultaneously

2 : =

= 43.75 sec 44 sec

1
= 6.25
0.16

Create a plot by clicking the XY Plot button, which should


be visible in both the button bar below the menus and in the
Graph pallet. Get into the habit of entering the independent
variable, t, and its limits, before entering the function name
or expression as the dependent variable. Reversing the order
leads to Mathcad trying to be helpful and autoscaling using
its standard range of 10 to 10, which is rarely the range we
will want. Format the plot, adding gridlines, axes labels, and
a plot title, Fig. A2.8.
We can reuse function and variable names. The assignment operator := is a local assignment, meaning that it can
be overwritten by a new assignment operator which appears
to its left or below it in the worksheet. If we reuse a vari-

Space Bar Space Bar

to see

to see

v (t ) = 2.74e 0.16t cos ( 0.68t 2.91) + 2.66

v.2(t) : 2.74*e ^ 0.5* t

v.2( t)

0.013

10

tion by the Heaviside unit step function zeros out the value of
the response function, until the moment when the Heaviside
step function transitions from zero to one, Fig. A2.7.
As an example of a second-order oscillatory step response, we will plot the result of Sect.2.7.3

10

Fig. A2.7 The Mathcad plot of the product of the response function,
v(t), and the Heaviside step function, us(t). Mathcads notation for the
Heaviside step function is (t)

s1 , s2 = jw

*cos(0.68* t 2.91) + 2.66

able name, Mathcad underlines it with a green squiggle, to


alert us, in case we thought the variable name was unique.
Mathcad handles units as if they were variables, and has virtually every engineering unit predefined. Click on the measuring cup symbol to bring up the unit dialog box. Many
common choices for variable names are predefined units.
Consequently, variables may be underlined with a green
squiggle, even though they are unique, because they are also
the abbreviation of a unit.

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

110

Mathcad permits mixed units in calculations. Mathcad


converts all units to SI prior to computation, and then presents the results in SI (or sometimes metric) but adds a blank
placeholder. If the user enters a unit into the placeholder,
Mathcad recomputes and expresses the result those units.
For example, type in the following volume computation,
where the three lengths are expressed in inches, feet, and
centimeters.
5*in *0.6*ft *14*cm =
You will see,
5 in 0.6 ft 14 cm = 3.252 L
Edit the place holder. Type in^3. You will see
5 in 0.6 ft 14 cm = 198.425 in 3

Plotting in MATLAB
MATLABs Environment
MATLAB is a programming environment. The default
configuration opens with a tabbed tool bar across the top of
the screen and five windows below it, with the Command
Window in the center. MATLAB is an interpreter which
means MATLAB translates and executes code line by line.
The practical effect is that the command line, identified
by prompt>>in the Command Window, can be used as a
calculator. For example, typing 2+4 Enter at the command
prompt yields
>>2+4
ans =
6
MATLAB keeps a record of the commands and the variables
used in calculations and scripts. The Command History window shows a history of commands entered which extends
to prior uses of MATLAB. It is convenient for an individual
using MATLAB on a personal machine, but it also means
that users of a public machine have access to the prior users
command history. The command history can be cleared by
right clicking on the title bar at the top of the Command History window to bring up a context menu and selecting the
item Clear Command History.

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

the MATLABs default variable type is an array. An array


is an ordered set of data where an individual element is identified by an index. The indices locate elements in an array
as if they were Cartesian coordinates. A one-dimensional
array, also called a vector, because it resembles mathematical vectors, is a sequence where an element is identified by
a single index. Two-dimensional arrays, which resemble matrices, require two indices to identify an element. In standard
mathematical notation, the index of an element in a vector or
matrix would be subscripted. There are no subscripts, superscripts, or Greek characters in computer code, only the keyboard characters. Array indices are contained in parentheses.
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
values of vector indices must be positive integers. A restriction of MATLAB is that the smallest vector index is one, not
zero.

Assignment Statement
The assignment operator in MATLAB is a conventional
equal sign,=. Assignment statements are read from right to
left. The right side can be a numerical value, a previously
defined variable, or an expression. The left side must be a
variable.
For-End Loop
MATLAB provides a number of flow-control instructions,
or commands, which allow the execution sequence of repeat
or skip blocks of code. We will use a logic structure called
a for-end loop which, as one might guess, begins with the
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
is defined in the for line and given an integer range. The
counter variable consecutively assumes values of a defined
range. The lines of codebetween the for and end lines are
repeatedly executed, until each counter variable defined in
the for line and incremented by one passes through, and the
loop reaches its upper limit.
Example. We will wish to repeat the calculation of a response function, perhaps one thousand times. We choose a
name, say n, for the counter variable. We can use the counter
variable in the code between the for and end statements, but
we must not change its value. MATLAB will increment the
value of the counter variable, and check it against the upper
limit of the range. The syntax of a for-end loop that executes
10 times is
for n=1:10
instruction;
instruction;
instruction;
end

Mathcad and MATLAB

We can create a vector variable and assign it values using a


for-end loop. Say we need a time vector t from t=0 to t=1.0
with the time step or increment t = 0.1. How many elements are in the vector t? There are 11. The common error is
to divide the interval by the increment,
tend tstart 1.0 0
=
= 10 Intervals
t
0.1
This is the correct number of intervals of one-tenth between
zero and one, but we are short of one value needed to create
the end point for the last interval. The calculation of interval
end points is
tend tstart
1.0 0
+1 =
+ 1 = 11 Interval End Points
t
0.1

MATLABs Editor
We will use MATLABs editor to write a script or program
to create the vector t., stored with the extension.m, and then
run at the command line by typing the files name. Search as
you may, you will not find a tab or window for MATLABs
editor from the default environment. The editor is opened,
when you choose the icon, New Script, in the Home tool
strip tab. The Editor opens as a second program. The only
way to navigate from the Editor back to the MATLAB environment is by clicking on the MATLAB icons in the Windows task bar, and selecting the M
ATLAB environment.
Comments are labels, explanations, and notes added to
computer code for the programmers and future users benefit. The longer and more complicated the code, the more
important are comments for structuring the program and
making it understandable. Comments begin with a percent
sign % and are colored green in MATLABs editor. A comment may occupy the end of a line of code, or may be a line
by itself. Comments which identify variables and describe
the function of blocks of code are essential. The purpose of
writing a script is to save time by automating tasks. Poorly
commented code is difficult to understand and use, and defeats the purpose.
We will begin our script with a comment giving the name
of the script, First.m, and the date. MATLAB names cannot have spaces. Use an underscore instead. Our simple programis
% First.m 06-02-14
for n=1:11
t(n)=0.1*(n-1)
end
Write and save the script. Although you can accept MATLABs default location, it is best to save your script to a per-

111

sonal storage device or location. Click on the green triangular Run icon in the tool bar at the top. A dialog will appear
stating that the script is not in the current folder or on the
MATLAB path. Click on the button, Add to Path, which adds
your personal storage location to those locations that MATLAB checks for scripts. It will also run the script. Nothing
appears to happen, because you are still in the Editor. Navigate to the MATLAB environment. You will see the result in
the Command Window. The elements of vector t are written
to the Command Window for each iteration through the loop.
Notice that t is now listed in the Workspace window, and
has the Value,<111 double>, describing it as a one-dimensional array with 11 elements of a data-type double. Data
type refers to how data are stored. Double refers to double
precision floating point number. Double precision uses 64
bits of computer memory. Floating point refers to scientific
notation.

Plotting
We will now edit First.m to create a second vector y and
plot it. Although MATLAB has pi as a constant, it does not
have Eulers number e. Exponentiation to the base e is performed with the function exp(). MATLABs function, plot(),
produces an interactive plot which can be resized, formatted,
and interrogated using the data cursor. The arguments of
plot() are pairs of vectors which contain the x and y-axes coordinates of the data plotted, e.g., plot(x, y). In our example,
the independent variable is the vector t. Add the plot command after the end statement.
Our output is now graphical in the form of a plot. We do
not need to see the values of t and y written to the screen. In
fact, writing each iteration of the loop to the screen slows
execution of the script significantly. A semicolon at the end
of line of code is an instruction not to write that line on the
screen.
% First.m 06-02-14
for n=1:11;
t(n)=0.1*(n-1);
y(n)=exp(-t(n));
end;
plot(t,y)
Save and run this script. The MATLAB plot titled Figure1
will appear as a new document, Fig A2.9. If Figure1 does
not exist, it will be displayed on top of the Editor. If Figure1
already exists, it will be overwritten with the current plot but
will not be brought to the front. Navigate to it by clicking
on the MATLAB icons in the task bar at the bottom of the
screen.
MATLAB figures can be formatted after they are created,
by opening the Tools menu item in the menu bar of the
figure, and choosing Edit Plot. It is also possible to format

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

112
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

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

Fig. A2.9 Plot of the MATLAB script, First.m

figures with commands in the script. Although more complicated, it allows the formatting to be automated. The technique is addressed in the Programming in MATLAB appendix to Chap.8.

The Dark Side of MATLABs Workspace


There is a dark side to MATLABs workspace. Variables remain in the workspace, until they are deleted, by selecting
the variable in the Workspace window, right-clicking, and
choosing delete, or the entire workspace is cleared by selecting the title bar of the Workspace window, right clicking, and
choosing Clear Workspace. The existence of variables in the
workspace can mask errors and omissions in scripts. For example, say you write a script to calculate a response function.
You intend to define a variable named tau equal to a certain
value, but you simply forget and omit that line of code. If you
run the script and tau was not used in any previous calculations and, consequently, is not in the workspace, MATLAB
identifies the error. A ding sounds, and the Command windows displays an error message indicating that an undefined
variable was used in line x of the script. On the other hand,
if you omit the definition of tau in the script, but tau exists in
the workspace with a different value, MATLAB will use the
workspace value. The script will execute but produce an erroneous result.
Minor editing of the script, First.m, will provide a
graphical example of the effect of MATLABs workspace.
Second.m uses the same variables, named t and y, as First.m.
The upper limit of the for-end loop is reduced to 9 from 11.
Finally, the argument of exp() is multiplied by two, increasing the exponential decay rate.
% Second.m 06-02-14
for n=1:9;
t(n)=0.1*(n-1);
y(n)=exp(-2*t(n));
end;
plot(t,y)

0.2

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

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

The plot, Fig. A2.10, is correct from t=0 to t=0.8, corresponding to the portions of the vectors, t and y, which were
overwritten with new values. The final two elements of vectors t and y were not overwritten, and still hold values created by the previous script First.m, leading to the kink in the
curve. Although this example may seem contrived, it is not.
Engineers tend to use and reuse the same meaningful variable names. MATLAB does not create a new instance of an
existing variable. Consequently, when MATLAB executes
an error, or the results of successful execution of a script do
not make sense, clear the workspace and run the script again.

Plotting Flow Chart and Script


We now have the basics to write a MATLAB script to plot
response functions. We must develop the logic to establish
the duration of the plot and the time step of the calculation.
The variable is tau_max, the largest time constant of the
system. tchar is the smallest characteristic time of the system, which may be either a time constant, or, if the response
is oscillatory, a period of oscillation T, Sect.2.9.4. The time
step, dt, is 1/200 of tchar. The duration of the calculation,
tend, should be approximately six or seven tau_max. The
number of calculations for the plot N is
N=

tend
dt

The for-end loop is similar, and the plot() command is identical to scripts, First.m and Second.m. Figs. A2.11 and A2.12
are flow charts of the script.
We cannot name the script, plot.m, because plot () is a
MATLAB function, so we will name it Plotting.m. Because
we must edit the script each time we wish to plot a new response function, it is helpful to uniquely identify each version 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
script, Plotting.m

Start

Calculate maximum time constant from the


minimum magnitude real eigenvalue component

sigma_min = mininimum ||
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
maximum magnitude real eignevalue component

sigma_max = maximum ||
tchar = 1/abs(sigma_max)

Calculate the characteristic time tchar as the


period of the highest frequency oscillation

omega =
tchar = (2*pi)/omega

Calculate the time step dt

dt = tchar/200

Continued in Fig. A2.12

% plotting_060214A.m
% Response Function Plotting Script 6-2-14
%
% tau_max is the largest time constant in the system.
% If the eigenvalues are known then use the smallest
% magnitude real component sigma.

% system. If the smallest characteristic time is a


% period, use the following block to calculate the
% period from an angular frequency in radians/sec.

sigma= -0.5
tau_max = abs(1/sigma)

% If the smallest characteristic time is a time


% 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.


tend = 6*tau_max
% tchar is the smallest characteristic time of the

% omega = 0.68
% tchar = 2*pi/omega

tchar = tau_min
% dt is the time step between calculations. Use 1/200
% of the smallest characteristic time.

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

114

Continued from Fig. A2.11

v, m/sec

Calculate the number of interations of for-end loop

Response v(t)

N = tend/dt

0
Initialize for-end loop counter variable

n=1

t, sec

10

12

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

Compute output vector element

y(n) = response(t(n))

% plotting_060214B.m
% Response Function Plotting Script 6-2-14
%
% tau_max is the largest time constant in the system.
% If the eigenvalues are known then use the smallest
% magnitude real component sigma.
sigma = -0.16
tau_max = abs(1/sigma)

Increment counter by one

No

n == N ?

n=n+1

Yes

Plot vector pairs

plot(t,y)

Finish

Fig. A2.12 Lower half of the flowchart of the MATLAB script,


plotting.m

dt = tchar/200
% Calculate the upper limit of the for-end loop. This
% value is the number of calculations.
N = tend/dt
for n=1:N
t(n)=(n-1)*dt;
y(n) = 2.66*(1-exp(-0.5*t(n)));
end;

% tend is the plot duration. Should be 6 or 7 tau_max.


tend = 6*tau_max
% tchar is the smallest characteristic time of the
% system. If the smallest characteristic time is a
% period, use the following block to calculate the
% period from an angular frequency in radians/sec.
omega = 0.68
tchar = 2*pi/omega
% If the smallest characteristic time is a time
% constant then comment-out the two lines above
and use
% the following line.
%tchar = TimeConstant
% dt is the time step between calculations. Use 1/200
% of the smallest characteristic time.
dt = tchar/200
% Calculate the number of calculations N. This is the
% upper limit of the for-end loop.

References and Suggested Reading

115

Response v2(t)

4.5

3
2
1

v, m/sec

v2(t), m/sec

Response Plot with Homogeneous and Particular Solutions

1
1
2
0

10

20

t, sec

30

40

Fig. A2.14 The MATLAB plot of the response function, v2(t), from the
script, plotting_060214B.m, after formatting

N = tend/dt
for n=1:N
t(n)=(n-1)*dt;
y(n)=2.74*exp(-0.16*t(n))*cos(0.68*t(n)-2.91)+2.66;
end;
plot(t,y)
Plotting additional traces on the y-axis requires the creation of additional output vectors, and then adding pairs of
independent, dependent variable vectors to the plot statement. We will rename the script, plotting_060314A.m,
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
axis. The revisions are in the for-end loop and in the command plot (). The resulting plot is Fig. A2.14.
% plotting_060214C.m
% Response Function Plotting Script 6-2-14
%
% tau_max is the largest time constant in the system.
% If the eigenvalues are known then use the smallest
% magnitude real component sigma.
sigma= -0.5
tau_max = abs(1/sigma)
% tend is the plot duration. Should be 6 or 7 tau_max.
tend = 6*tau_max
% tchar is the smallest characteristic time of the
% system. If the smallest characteristic time is a
% period, use the following block to calculate the
% period from an angular frequency in radians/sec.
% omega = 0.68
% tchar = 2*pi/omega

t, sec

10

12

Fig. A2.15 The MATLAB plot of the response function, v1(t), 2.660.5t
and 2.66, from the script, plotting_060214C.m, after formatting

% If the smallest characteristic time is a time


% constant then comment-out the two lines above
% and use the following line.
2 tchar = tau_min
% dt is the time step between calculations. Use 1/200
% of the smallest characteristic time.
dt = tchar/200
% Calculate the upper limit of the for-end loop. This
% value is the number of calculations.
N = tend/dt
for n=1:N
t(n)=(n-1)*dt;
y1(n) = 2.66*(1-exp(-0.5*t(n)));
y2(n) = 2.66*(exp(-0.5*t(n)));
y3(n) = 2.66;
end;
plot(t,y1,t,y2,t,y3)

References and Suggested Reading


Hildebrand FB (1976) Advanced calculus for applications, 2nd edn.
Prentice-Hall, Englewood Cliffs
Ogata K (2003) System dynamics, 4th edn. Prentice-Hall, Englewood
Cliffs
Ogata K (2009) Modern control engineering, 5th edn. Prentice-Hall,
Englewood Cliffs
Rowell D, Wormley DN (1997) System dynamics: an introduction.
Prentice-Hall, Upper Saddle River
Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
dynamics. Addison-Wesley, Reading

Introduction to the Linear Graph Method,


Step Responses, and Superposition

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 systems 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 systems 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.

3.1Introduction
We revisited differential equations in Chap.2 and thoroughly
reviewed the method of undetermined coefficients. In this
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 engineering design process. Specifically, our approach will allow us
to make relatively quick decisions using simple models, before we commit ourselves to the time and expense of a more
accurate but involved mathematical model. We will focus on
the relationship between the input to the system and the response of the system. From this perspective, the differential
system equation is seen as an operator, where the system
equation operates on the input variable to yield the output
variable, Fig.3.1. Although we use the term input variable
in order to predict the behavior of a physical system, we need
to know how the input variable changes with time. In other
words, we need the input function, also known as the forcing
function, to calculate the response function.
This perspective is attractive for three reasons. First, it
simplifies the solution of the differential system equations in
many cases. Second, it allows the solution of impulse, pulse,

step, and sinusoidal inputs to be extended to arbitrary inputs


by using superposition. Lastly, and most importantly, it allows much of the content of this course to be used in engineering practice.

3.2 Introduction to the Linear Graph Method


Chapter1 introduced the definitions of mechanical work and
power, the concepts of mechanical and energetic models,
and the techniques and notation needed to describe power
flow through an energetic network. We will now develop
these basics into the linear graph method. A linear graph is
a circuit-like representation of an energetic system drawn
from a schematic representation of an energetic model. The
technique of drawing a linear graph from the schematic of an
energetic system is straight forward. We shall see in Chaps.4
and 5 that the same methodology is applied for mechanical,
fluid, thermal, and electrical systems. The greatest benefit
of a linear graph is the ease of joining dissimilar subsystems
into a single energetic system which will allow us to model
machines, as we shall see in Chap.6.

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_3, Springer Science+Business Media New York 2014

117

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

118

Dynamic
System

Input Variable
Forcing Function

Response Function

Fig. 3.1 A dynamic system operates on an input yielding a response.


The equation which describes the response is the response function

x,v
F(t)

Mass

x,v

Output Variable

Lubricating
fluid film
Damping b

Fig. 3.2 Schematic energetic model of a force source, F(t), sliding a


mass M on a lubricating film with damping b

We will introduce the linear graph method with the example of a force source sliding a mass on a lubricating fluid
film, Fig.3.2. The lubricant film supports the mass to prevent
it from making solid to solid contact with the surface. In order
for the mass to slide, the fluid must shear, which dissipates
mechanical power, thus heating the fluid. This phenomenon of
viscous friction is called damping, and is represented by the
parameter b. We will present the basics of the viscous shear of
fluids, as we examine the energetic properties of the model.

3.2.1 Energetic Model


First, it is important to note that an energetic model need not
resemble the physical object in the slightest. The energetic
model shown in Fig.3.2 can represent any number of physical systems of vastly different appearance, size, and function, such as a fuel injector for an automobile or a massive
table of numerically controlled machine tools. Appearance
is not a physical attribute we wish to represent in our model,
except in those cases, when the geometry of a device affects
its energetic behavior. The size of a system is conveyed by
the magnitudes of the systems energetic parameters, M and
b, which need not be established, until the model is actually
used to predict the response of a specific physical system.
An energetic model is an abstract model. Engineering
students are intimidated by the term, abstract, since it is associated with intellectually challenging aspects of mathematics
and physics which would more correctly be titled, arcane,
rather than abstract. A more common usage of abstract denotes greatly simplified. The abstract of this chapter is a
summary of the main points.
A difficult aspect of modeling for most students is the
fact that an element in an energetic model represents a single

Surface 1
v1
h

Fluid Shear
v2
Surface 2

Fig. 3.3 Shear of a fluid between two planar surfaces separated by


distance, h, moving with horizontal velocities, v1 and v2

e nergetic attribute of the physical system. In physical reality,


an object with mass has additional properties or attributes,
such as volume, compliance, etc. All physically real objects
have multiple energetic attributes. However, an element
in an energetic model has just one energetic attribute. The
consequences of a model element having a single energetic
property are contrary to our everyday experience with real
physical objects. A mass element in an energetic model can
only store kinetic energy. It cannot store strain energy nor
can it dissipate energy as heat through inelastic deformation.
Consequently, a mass element is ideally rigid. It cannot deform. Likewise, a fluid which shears and dissipates energy as
heat can only dissipate energy. It cannot store energy. Consequently, it has no mass and is incompressible.
If a physical object has two significant energetic attributes, then two energetic elements are needed to model the
one physical object. Judging the significance of various energetic properties of a single physical object or system must
be deferred for now, since it requires interpretation of the results of the modeling process. As a rule of thumb, for the first
and simplest model, an attribute is significant and should
be included in the model if it stores or dissipates at least 10%
as much energy or power as any other similar element in
the model. Any model should follow the KISS principle,
where KISS stands for Keep It Simple, Stupid. Modeling
should always progress from simple to more complicated.
It is possible to add elements to a model, if the preliminary
results indicate it is necessary.
We shall see that energy dissipation in mechanical systems is difficult to describe mathematically. The mass of
a solid object can be represented within the accuracy with
which we can determine the objects mass. Conversely, all
descriptions of friction are at best approximations of the
actual energy dissipation, and are invariably uncertain to a
greater degree than other energetic properties.
The fluid film provides lubrication, that is, supports the
mass preventing contact between the two solid surfaces. For
the mass to slide horizontally, the fluid must shear, Fig.3.3,
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


of the gap, h, between the mass and the supporting surface
varies with the velocity of the mass, increasing with increasing translational speed. Also, there is a minimum velocity,
below which the asperities (or peaks) on the solid surface
of the mass would contact the asperities on the supporting
surface, creating Coulomb friction. We will model the system as having a constant gap height, h. The value we would
use for h would be that which we believe is representative
of our systems operating condition. In other words, a gap
height should be typical of that which occurs, when the mass
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
sticks to that surface. This is known as the no-slip assumption, and results in the fluid in contact with a given
surface having the velocity of that surface. Macroscopic
shear displacement deforms the fluid through shear strain,
, which is given as

x
=
(3.1)
y
where x is the difference in the horizontal displacement of
parallel surfaces of a cube of fluid, and y is the length of the
side of the cube and, thus, the distance between the cubes
parallel faces, Fig.3.4.
The shear strain rate is the time derivative of shear strain:
d
1 x 1 dx
= lim
=
(3.2)
dt t 0 t y y dt
If the shear strain is uniform across the height h of the gap,
then dx / dt = v and the shear strain rate is
d v
=
(3.3)
dt
y
The shear stresses within the fluid, Fig.3.4, are related to the
strain rate of the fluid by the parameter viscosity given the
symbol, . The viscosity of most common liquids is a function of strain rate and temperature. Therefore, in the general
case, the relationship between shear stress and strain rate of
a liquid is


d d
,T
dt dt

(3.4)

Equation3.4 is non-linear, because the viscosityis a function of the shear rate, d / dt . It is also not stationary, meaning that the equation does not remain the same. In this case,
it changes with temperature. If fluid temperature was held
constant, or the viscosity was insensitive to (did not vary

b x

yx
y

xy

y
x

xy

yx
Fig. 3.4a Shear stresses, xy and yx, on the surfaces of a cubic element
of fluid. bShear in the xy plane of a fluid cube

much with) temperature, Eq.3.4 could be rewritten with the


shear rate raised to the power n:


d
dt

= 0

(3.5)

where 0 and n are constants. This is a common form of a


non-linear relationship between shear rate and shear stress
in fluids.
When viscosity changes little over the anticipated ranges of shear rate and temperature which the fluid will experience, then viscosity can be modeled as an ideal, constant
viscosity. The shear rate and shear stress are then linearly
related, Eq.3.6,
d
=
(3.6)
dt
A fluid modeled as having a constant viscosity is known as
a Newtonian fluid, and yields the linear velocity distribution
shown in Fig.3.3. We will assume an ideal, constant viscosity, that is, Newtonian fluid, as the lubricant, which is the
simplest model of the fluid mechanical effects.
The free body diagram of the forces acting on the mass is
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,
F(t), acts in the positive direction, then the viscous friction
shear force, Fshear, acts in the negative direction. The magnitude of the shear force, Fshear, is integral of the shear stress,
yx, over the bottom area of the mass. Similarly, integration
of the fluid pressure across the bottom of the mass yields
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
rotates the mass about the z-axis. This rotation cannot happen without violating the geometry we have assumed for
our model. What we have neglected is the variation of gap
height h and the pressure acting on the mass with position.
The variation in h and pressure would be the result of a more
rigorous fluid mechanics model of the motion of the mass on
a fluid film with constant viscosity. A fluid mechanics model
would reveal that, if the mass moved fast enough, the rota-

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

120

F(t)
y

FR

FShear

Fig. 3.5 Free body diagram of the forces acting on mass M

x,v
F(t)

Mass
M
FShear

Fig. 3.6 Free body diagram showing only the forces acting on mass M
that do work

tion would be clockwise, opposite of the couple shown on


the free body diagram, leading to a gap height that decreased
from the leading side to the trailing side of the mass. We
are constructing, however, an energetic model of the system.
We need not include these details, unless they have a significant effect on the energetic behavior of the system. We
will neglect the rotation of the mass. We will examine how
to couple translational and rotational motions in mechanical
systems when we develop transducers in Chap.6.
We can further simplify the free body diagram by eliminating forces or force components which do no work on the
system. In this model, motion is restricted to the x-axis. The
weight W of mass M and reaction force FR created by the
fluid pressure act in the y-direction. They have no component force in the x-direction and, consequently, they do no
work on the system. We can eliminate them from the free
body diagram of the model, yielding the simplified free body
diagram shown in Fig.3.6.
Although we can dismiss the details of the results of fluid
mechanics, which contradict the geometric constraints of our
simple model, we cannot be as cavalier with any energetic effects, since we are creating an energetic model. The fluid lubricant in the model, shown in Fig.3.2, has a single energetic
property, damping; it dissipates the mass energy of motion
as heat in the fluid. If the heat dissipated remains in the fluid
and enough heat is dissipated to raise the fluids temperature
and decrease the fluids viscosity, making it easier to shear,
then the damping property or coefficient b of the fluid would
change with motion of the mass. Energetic parameters, such
as the fluids damping coefficient b, must be constant if the
system model is to yield a linear differential equation.

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


removing either the heated fluid, or just the heat, then the
temperature increase will be minimized. If heat accumulates
in the fluid, then the temperature increase will be maximized.
Hence, for a given fluid, the magnitude of the effect of the
energy that is dissipated in the fluid on the fluids viscosity
depends on these following factors: the thermal properties
of the fluid and the solid surfaces; whether it is an open
or closed fluid system (does the fluid recirculate or does
new fluid enter and heated fluid leave?); and the operating
conditions. Thermal systems and heat transfer are discussed
in Chap.5.
There are two ways to include a more accurate model
of viscosity in our energetic model. We can maintain a
linear model by approximating a non-linear function. The
method, known as linearizing about an operating point,
is developed in Sect.4.2.4.1, when we revisit damping in
Chap.4. The second approach is to formulate a more accurate non-linear description of viscosity, such as Eq.3.4
or 3.5. Non-linear models yield non-linear differential system equations, which have a drawback, although they present more realistic descriptions of dynamic systems. Linear
system equations can be solved by a variety of classical
mathematical techniques. Conversely, non-linear differential system equations must be solved by numerical methods, which approximate differential equations as finite
differences. Numerical methods, discussed in Chap.8, are
widely used in engineering, but each analysis is its own
special case. The first model of any system should be a
simple linear model. If a non-linear model is found to be
necessary, then the simple linear model provides the starting point for the more complicated non-linear model. Since
we are developing a simple, first model of the mass sliding
on a lubricating fluid, we will ignore thermal effects. In
engineering jargon, the thermal effects on the viscosity of
the lubricant will be unmodeled dynamics.
Deciding to use a constant viscosity and a uniform gap
height h establishes the energetic model of shearing the lubricating film. The shear stress, yx, is described by Eq.3.6
and is uniform over area A of the bottom of the mass, since h
is constant. The shear force, Fshear, is
A dx A
(3.7)
Fshear =
=
v
h dt
h
where the velocity of mass v is also expressed as dx/dt.
The Newtonian formulation of the energetic model uses
dx/dt. The formulation derived using linear graph method
uses v.
Damping coefficient b is the parameter which relates the
velocity difference across the damping element, the fluid
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


across the fluid film,
A
A
Fshear =
v = bv b =
(3.8)
h
h

F (t ) = M

d2x
dx
+b
2
dt
dt

with the input variable, F(t), and the dependent or output


variable, x, on opposite sides of the equation. Rewriting the
derivative terms in the order of decreasing order, and then,
for standard form, clearing the coefficient of the highest-order
derivative of the output variable x yield the Newtonian form
of the system equation for the displacement of the mass:

1
d 2 x b dx
F (t ) = 2 +
M
M dt
dt

Collecting the effect of viscous friction, which is distributed across the bottom surface of the mass, into damping
coefficient b is an abstraction or simplification. A single
parameter that represents a property distributed across a
surface, such as viscous friction, or throughout a volume,
such as mass, is called a lumped parameter. This is a familiar approximation, since free body diagrams also lump
or localize properties. Collecting or concentrating properties which are actually distributed simplifies the resulting
differential system equation by eliminating the spatial variable, x. The system equation is then an ordinary differential
equation with time derivatives, rather than a partial differential equation with derivatives of both time and position.

System Equation

This form of system equation is a second-order differential equation. It can be rewritten in terms of velocity, using
the definition v = dx dt , to yield a first-order differential
equation:
1
dv b
F (t ) =
+ v System Equation
M
dt M
In Sect.3.5, we will introduce a special form for the firstorder system equations known as the time constant form,
where the output variable term is cleared of a coefficient,
bF (t ) =

M dv
+ v System Equation in Time Constant Form
b dt

Time
Constant

3.2.2Newtonian Formulation of the


Force-Mass-Damper Model
A Newtonian formulation of an energetic model is derived
in terms of the dependent variables, translational displacement x, and rotational displacement . A free body diagram is
drawn, showing the forces and torques acting on each mass,
and each mass moment of inertia. The sum of the forces acting on a mass accelerates the mass in translation. The sum
of the torques accelerates the mass mass moment of inertia.
We have restricted the motion in the mass-damper model to
translate in the x-direction, so we do not need to sum torques
about any axis or forces in the y or z directions. Referring
to the free body diagram, Fig.3.6, which shows only the xdirection forces which do the work, the sum of the forces is
expressed as
d2x
(3.9)
Fx = F (t ) Fshear = M dt 2
Substituting for the shear force yields
F (t ) b

dx
d2x
=M 2
dt
dt

This is not in the proper form for a differential equation. Recall from Chap.2 that the standard form is to separate terms

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
constant and given the symbol , Greek for t.

3.2.3Linear Graph Formulation of the


Force-Mass-Damper Model
A linear graph is a network which represents the energetic
properties of a dynamic system. A linear graph is drawn from
an energetic model usually represented as a schematic. The
schematic indicates both graphically and with call outs
(text) every significant energetic property in the dynamic
system. Each energetic property on the model is represented
as a single element in the network. The linear graph of a
mechanical system is drawn by identifying locations, called
nodes, in the system with distinct velocities, and then adding lines, called branches, between the nodes to represent
each energetic property. Each element in the circuit-like linear graph represents the power being sourced (provided),
stored, or dissipated in that element, at any instant in time.
We will begin with mass. The energetic property of a
mass is its storage of kinetic energy:


EM =

1
Mv 2
2

(1.34)

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

122

x,v
F(t)

Mass

Lubricating
fluid film
Damping b

g
Fig. 3.7 Schematic of the mass-damper system showing the velocity
nodes, node 1 on the mass, and node g on the ground

An ideal mass is rigid and has a single velocity. As you recall,


the velocity of a mass must be measured relative to a non-accelerating (inertial) reference. In practice, we make our measurements relative to the surface of the earth, which we will
call, not surprisingly, ground. We will likewise refer to the
frame or base of a machine as a ground reference, if it is
stationary relative to the earth. We will designate distinct velocities, such as the velocity of a mass and the velocity of its
ground reference, as nodes. The linear graph method uses
nodal subscripts to identify a velocity node. The subscripts
are usually numbers, except for reference nodes where abbreviations are used, that is, g for ground. The schematic
of the energetic model with the velocity nodes of the mass
and ground is provided in Fig.3.7.
The kinetic energy of the mass is written using the nodal
subscript notation, to identify the velocity difference between the mass and its inertial (ground) reference,
1
E M = Mv12g
(3.10)
2
Notice that the subscript on the kinetic energy E is M for
mass.
We now consider mechanical energy lost as heat in the
damping of the viscous fluid film due to shear. The strain
rate of the fluid was calculated as the difference between the
velocities of two parallel surfaces divided by gap height h.
The velocities of two surfaces were identified generically
as v1 and v2 in Fig.3.3. The surfaces are the bottom of the
mass and the top of the ground, to which we have assigned
the velocities, v1 and vg, respectively. The velocity difference
between nodes 1 and g, written as v1g, allows us to write the
force acting through the damping element, the lubricating
fluid film, in terms of damping coefficient b, and the velocity
difference across the damping element:
Fb = bv1g
(3.11)
Notice that the subscript on the force is the damping parameter, or coefficient b, rather than the subscript shear used
in Newtonian formulation.

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
force vector. The use of a vector indicates that the mechanism, or component, applying force to the system is either
unknown or unimportant to the analysis.
Recall that mechanical power is

P
= Fv
(1.17)
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
power source. A force source provides mechanical power to
the system at a known or specified force. The velocity of the
point of application is not known. The response of the system to the applied force F(t) determines the velocity of the
point of application of the force. Although force is a familiar
input to statics and dynamics problems, it is more common
in practice to know the velocity of a machine element, which
presses or pulls on a dynamic system, than to know the force
which is applied. A velocity source is a mechanical power
source, in which we know or specify the velocity of the point
of application of an unknown force. The force which the velocity source must provide is determined by the reaction of
the system to the velocity source.
All energetic elements in a linear graph require two velocity nodes, including the force source. It is convenient to draw a
force source as a vector, showing only the point of application
of the force. In reality, the force is applied by a mechanism
or mechanical element which needs to push or pull against a
suitable reaction in order to provide a force. Anyone who has
pushed or pulled, while standing on a slippery surface, can
attest to this fact. We will use the convention of an input force
shown as a vector with its point of application indicated, but
its tail floating in space, to indicate that it is reacting against
ground, since it must react against something.

3.2.3.1DAlemberts Force and Dynamic


Equilibrium
So far, the only differences between the linear graph method
and a Newtonian formulation have been minor changes in
notation and the use of velocity rather than position as the
dependent variable. The significant difference between the
two methods is the how force and torque equilibrium statements are written. Recall that the Newtonian formulation
began by equating the sum of the forces acting on a mass
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,
(3.12)
Fx = FNet = max
The essence of a Newtonian formulation is that the unbalanced forces accelerate a mass. Another way of stating this

3.2 Introduction to the Linear Graph Method

is in the inverse. When the forces acting on a mass are in


equilibrium, the mass does not accelerate. The equilibrium
requirement that the forces acting on a mass sum to zero,

= FNet = 0

is a static equilibrium.
A mass is fundamentally different from other energetic
mechanical elements. A single force can be applied to a mass.
In contrast, a spring must have equal and opposite forces at
both ends, in order to transmit or provide any force. You cannot push on just one end of a spring and develop a force.
There must be a reaction force pushing back against the other
end of the spring. The push back of a mass develops at the
point of application of the force since an inherent property
of mass is its resistance to acceleration. In 1743, DAlembert
introduced the concept of an inertial force, in the context
of extending the principle of virtual work from the static case
to a system of particles in motion. The DAlembert force,
also known as the inertial force, is the resistance of a mass to
acceleration. DAlemberts force is equal and opposite to the
net force, which accelerates a mass. An inertial force is the
pseudo-reaction force of a mass.
A dynamic equilibrium includes DAlemberts force, in
the sum of forces acting on the mass. A dynamic equilibrium is always zero, since, by definition, DAlemberts force,
max, equals the opposite of the Fx,
(3.13)
Fx max = 0
where DAlemberts force is
FD Alembertx = max
(3.14)
We will invert the sign of Eq.3.14 and use the opposite
of DAlemberts force as the force acting to accelerate the
mass, FM,
(3.15)
Fx FM = 0
Mathematically, the force acting to accelerate the mass, FM,
is the net force, Fx. Conceptually, FM=max. The distinction
between the two previous statements is the difference between a force, Fx, and the effect of that force, max.
There are two reasons for including the DAlembert
force (as FM) in the linear graph method. First, including
DAlemberts force allows a dynamic equilibrium statement to be written for every force summation. Second, by
naming the force acting to accelerate a mass FM, we have a
force associated with the mass element. We shall see that two
power variables per energetic element are needed to apply
the equivalent of Kirkchoffs circuit laws to a mechanical
system represented as an energetic network. The power

123

variables in a translational mechanical system are force and


velocity. Accordingly, we will express the acceleration in
FM = Max as the derivative of velocity,
dv
FM = M
(3.16)
dt

3.2.3.2 Through and Across Variables


Power is the product of two power variables in mechanical, fluid, and electrical systems. Only in thermal systems
is power a single variable. A set of analogies between the
power variables in different types of systems will allow us
to generalize and exploit the similarities between different
types of systems. There are two complementary sets of analogies in common use, one used with linear graphs, which
we are developing, and the other used with Bond Graphs.
Linear graphs resemble electric circuits, whereas the term
bond graph describes graphs which look like diagrams of
molecules with bonds between atoms.
The linear graph method categorizes power variables into
two classes, through variables and across variables. A
through variable acts or flows through an element. A through
variable has the same magnitude at both ends of an element,
with the exceptions of mass and mass moment of inertia and,
as we shall see later, fluid capacitance. An across variable
varies or drops across an element. There is a single value
of an across variable at a node. With reference to Sect.1.3,
where continuity and compatibility equations were introduced in the context of a piping system, volume flow rate
is a through variable. It is the same at both ends of a pipe.
Through variables sum to zero at nodes where elements connect. Pressure is an across variable. It decreases in the direction of fluid flow. Across variable differences sum to zero
around a closed-loop in a linear graph.
A useful method of identifying a power variable as a
through or an across variable is to consider how the quantity
is measured. For example, current must be routed through an
ammeter, or a fraction of the current routed through a parallel
shunt, for the current to be measured. Current is a through variable. Conversely, voltage cannot be measured with a single
probe at a single location. Voltage is a relative measurement.
We measure the difference in voltage between two nodes at
either end of an element or between a node at the end of an element and a reference node. In other words, we measure voltage across an element between the nodes connected at either
end or between (across) a node on the element and a reference
voltage. To measure a force, the load cell (force meter) must
be in the load path. The force must act through the load cell.
Force is a through variable. Conversely, the velocity of a component can be measured visually. Observation of the surface is
sufficient. However, velocity is measured relative to a reference velocity, usually, but not necessarily zero. The difference

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

124

Table 3.1 Linear graph and bond graph variable analogies


System Type
Power Variable
Linear Graph Bond Graph
Variable Type Variable Type
Fluid
Volume flow rate
Through
Flow
Pressure
Across
Effort
Electrical
Current
Through
Flow
Voltage
Across
Effort
Translational
Force
Through
Effort
Mechanical
Translational velocity Across
Flow
Rotational
Torque
Through
Effort
Mechanical
Rotational velocity
Across
Flow
Thermal
Heat flow rate
Through
Flow
Temperature
Across
Effort

between the components velocity and the reference velocity


is the value needed. Velocity is an across variable.
We will use the linear graph set of analogies to apply
Kirchhoffs current and voltage laws to the circuit-like network of our system. The linear graph and bond graph power
variable analogies are listed in Table3.1.

reference node. Since the mass is rigid, the two nodes on the
mass, identified as node 1, are a single velocity. Likewise,
two ground nodes identified as g are shown. They are also a
single velocity.
A linear graph is a network. It consists of a set of nodes
that represent distinct values of the across variable, connected by lines known as branches which represent the energetic elements. An arrowhead is drawn near the middle of
each branch, to define the positive direction for the through
variable in that energetic element. The direction defined as
positive for an element is arbitrary, except for sources and
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
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 derives its name from the use of lines identified by parameters,
rather than unique symbols for each type of element. An exception to the no-special-symbols rule is an input element or
source. The symbol for a source is a branch with the input
variable encircled in the middle of the branch.
Linear graphs are drawn by first drawing and identifying
(labeling) the nodes by a number or letter. In this example,
we draw and identify the velocity nodes, v1 and vg.
Next, draw the elements between their respective nodes,
which are identified on the schematic of the energetic system. Note the mass elements dashed line after the arrowhead. This convention indicates there is no reaction force
from the ground node against the mass. Both the force source
and the damper must have a reaction force at the ground, in
order to apply a force at node 1.

3.2.3.3Drawing a Linear Graph from a Mechanical


Schematic
The first and the most important step in drawing a linear
graph is to find and label nodes of distinct values of the
across variable on the schematic of the energetic system.
The across variable in mechanical systems is velocity. We
cannot see most of the variables we work with, but we can
see motion. Distinct values of velocity are those at either
end of the elements in the schematic of an energetic system,
with the exceptions of a mass, a mass moment of inertia, and
a fluid capacitance. A mass and a mass moment of inertia
are rigid. They have a single velocity. The second velocity
node associated with them is the velocity of the inertial reference. Any non-accelerating reference can be used as the inertial reference velocity. We will refer to an inertial reference
which is stationary in the local frame of reference as ground.
Figure3.8 shows nodes on either side of the vector, which
represents the force source, as well as nodes on either side of
the fluid film, which contributes to the damping b. The latter
two nodes also serve as a node on the mass and an inertial

3.2.3.4 Power Flow in a Linear Graph


Sources power the systems, and they are classed as active elements. The positive direction for the through variable of a source element is the direction of an increase in
the across variable. In our system, the force source reacts
against ground which has zero velocity. The point of application of the force will move and, thus, has non-zero velocity.
A positive force increases the velocity of node 1, relative to
ground velocity, and a negative force decreases the velocity
of node 1, relative to ground. The arrow on the force source
points in the sources positive direction, from the ground
node to node 1. Elements which store or dissipate energy are
classified as passive elements. The mass and damper are
passive elements. The positive direction for the through variable of a passive element is the direction of a decrease of the
across variable. Consequently, the arrows on the mass and
the damper elements point from node 1 to ground.
Each element in a linear graph represents a power flow. An
elements power flow is the product of that elements through
variable and the difference between the across variables on

x,v

F(t)

Mass
1 M
1

Lubricating
fluid film
Damping b

g
Fig. 3.8 Schematic of the mass-damper system, showing a reaction
against ground for the input force source

3.2 Introduction to the Linear Graph Method

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


the arrowhead, specifies the order of the across variable difference, 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 either stored or dissipated. The power flow of a source is usually
out of the source and into the system, but ideal, linear power
sources can also sink or accept power from the system.

3.2.3.5Node Equations, Equilibrium,


and Continuity
The network, or circuit-like linear graph representation of
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 Kirchhoffs current law,
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
control volume is positive. A through variable directed out of
a node or control volume is negative.
The through variable in a translational mechanical system is force. Consequently, a node equation is a statement of
dynamic equilibrium, which is possible by including in the
summation the force acting to accelerate the mass, FM. Summing through variables at node 1, the force from the source,
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
accelerate the mass, FM, are directed out of node 1 and are
negative,
F (t ) Fb FM = 0
(3.17)
Notice that we have written one node equation less than we
have nodes in the linear graph. We will always write one
node equation less than we have nodes. We will always omit
the node equation for the ground node. The omission of a
node equation for the ground node will prevent the inclusion
of a dependent equation within the set of energetic equations
to be reduced to a differential system equation. It will also
avoid the use of a fictitious reaction from ground for masses
and mass moment of inertias.

3.2.3.6 Path and Loop Equations, Compatibility


The across variable in a translational mechanical system is
velocity. The second set of equations, extracted from the linear graph, is written by summing across variable differences,
either between two nodes along different paths or around a
closed loop. These equations are analogous to Kirchhoffs
voltage law.
We will identify the across variable difference between
the ends of an element by subscripts which denote the nodes
the element is connected between. We will not denote the

125

F(t)

Fig. 3.9a Nodes of distinct values of the across variable velocity.


b Linear graph of a force source driving as mass-damper system

F(t)

F(t)

Fig. 3.10a Control volume cutting branches and through variables at


node 1. b Control volume at node g

element itself. This may seem awkward at first, but it reduces


the number of equations we need to work with. Again, for
passive elements, the across variable decreases or drops in
the positive direction of the through variable, as defined by
the arrowhead on the branch. The difference in the across
variable is calculated in that order. Referring to Figs.3.9 and
3.10, the positive direction of the force acting through the
damper, Fb, is from node 1 to node g. Hence the velocity
drop across the damping element is as follows:
v1 vg v1g
(3.18)
Note the use of the node number or letter as the subscript,
to indicate the velocity of a single node. The difference between the velocities of the nodes, 1 and g, is indicated by the
use of both subscripts in the order in which the difference
was calculated.
The use of node subscripts to identify the across variable
difference is contrary to the notation commonly used for the
voltage drop in an electric circuit, which is identified by the
elements subscript. Using the electrical engineering convention, the velocity difference across the damping element
is expressed as
v1 vg vb
(3.19)

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

126
Fig. 3.11 The mass-damper
loop is from node 1 across the
mass to node g, then back across
the damper to node1. The massdamper paths are from node 1
to node 2, across the mass and
across the damper

positive and negative differences for the values to be compatible and sum to zero. The loop equation for the massdamper loop of the system from node 1 to node 1 is

v1g v1g = 0
M (3.20)

Equation3.20 is written with the across variable differences


in the positive direction, as defined by the arrowheads on
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 subscript notation is simpler, the advantage of the node subscript
notation is clear, when there are elements in parallel, as in the
force source, mass, damper system, Fig.3.10:
Velocity
Difference
Across the
force source
from node
1 to node g
Across the
damper
from node
1 to node g
Across the
mass from node
1 to node g

Across Variable Drop

Element Subscript Notation

Node Subscript
Notation

v1 vg

vsource

v1g

v1 vg

vb

v1g

v1 vg

vM

v1g

The node subscript notation, which we shall use, makes it


obvious that all of the elements have the same velocity difference or drop across them. The element subscript notation
requires two equations to equate the velocity drops:
vsource = vb and vb = vM
The maximum number of unique loop or path equations,
which can be formed in a circuit or a linear graph, equals the
number of holes between elements in the network. This
maximum is not guaranteed, however, because the number
of elements which form the holes is a second factor. We have
two holes in this network. In the general case, we could form
at most two unique loop or path equations. As it happens,
with this particular network, we can only form one unique
loop or path equation. Further, that lone equation is a trivial
equation, meaning it contains no useful information. We will
write it for the sake of completeness, but it cannot be used in
the derivation of the differential system equation.
Loop equations and path equations contain the same information, Fig. 3.11. They differ only in signs. A loop equation equates the sum of the across variable differences between nodes around a loop to zero. Since a loop equation
begins and ends at the same node, the sum must contain both

v1g = (v1 vg ) = v1 + vg = vg1


(3.21)
This notation is not recommended. Reversing the subscripts
on the across variable hides the negative sign, and tends
to lead to subsequent sign errors, when the subscripts are
inadvertently written in the positive order. It is best to use
a negative sign, rather than to reverse the order of the node
subscripts to invert the sign.
A path equation equates the across variable differences
along two different paths between the same two nodes and
does not need to contain negative values. The path equation from node 1 to node g, across the mass and across the
damper, is
v1g = v1g
(3.22)
Equation3.22 is a trivial equation. It is correct but without
any useful information.

3.2.3.7 Elemental or Constitutive Equations


The continuity and compatibility equations describe how the
elements of a linear graph are connected. The elemental or
constitutive equations relate the two power variables of an
element to each other. The elemental equations for our system describe the forcevelocity relationships in the mass and
the damper.
The familiar equation, F=Ma, when written as an expression of the systems power variable, velocity, becomes the
general form of the element equation for mass,
F=M

dv
dt

From the general form, we write the element equation for the
systems specific mass element, by adding the identifying
subscripts to the force, and acting to accelerate the mass, FM,
as the velocity drops across the mass, v1g,
dv1g
(3.23)
FM = M
dt

3.2 Introduction to the Linear Graph Method

Fb

x 1 ,v1
Node 1

x 2 ,v 2

127

Fb

Node 2

Fig. 3.12 A dashpot schematic symbol for translational damping.


Compression, or extension, of the pistoncylinder forces fluid through
a flow restriction, shearing the fluid

We developed the relationship between the force, shearing


the viscous lubrication film, and the velocity of the mass,
Eq.3.11, when we derived the Newtonian formulation,
Fb = bv1g
(3.11)
Although Eq.3.11 was developed for the specific case of
shearing a viscous film between two planar surfaces, it is
the form of the elemental equation for all linear translational
damping. A mechanical schematic symbol to represent translational damping is a cross section of a piston inside a cylinder, Fig.3.12. The physical device is called a dashpot
and dissipates mechanical power has heat by pumping a fluid
through an orifice, a small diameter tube, or a small annular
space. The working fluid is typically oil but air is also used
in some designs.
Although we can write a functional description of the
input variable, F(t), we cannot write an elemental equation
that relates the through variable, F(t), to the across variable,
v1g(t), of the source. The velocity across the source will be
determined by the dynamic response of the system to the
force, F(t), imposed on the system. It is physically impossible to specify both power variables of the source.

3.2.3.8 Energy Storage Equations


The fourth category of energetic equations expresses the
amount of energy stored in individual energy storage elements, as well as in the system as a whole. Recall that we
are neglecting the energy dissipated as heat by shearing the
lubricating fluid. The energy stored in the mass-damping
system is stored in the mass as kinetic energy. The familiar
equation for kinetic energy is written in terms of mass parameter M and velocity relative to ground, v1g,
1
E M = Mv12g
(3.24)
2
Equating the energy in the system with the energy in the
mass,
E sys = E M
(3.25)

These are very important equations, but they are not used to
derive the system equation. They are used to establish the
initial condition needed to solve the equation, as we shall
see. The order of a differential system equation, which re-

lates one of the power variables in the system to the input


variable, equals the number of independent energy storages in the system. If there is only one energy storage element (in this case, mass), that energy storage element must
be independent. Consequently, the differential system equations derived below will be of first-order.
The energy storage variable, v1g in this system, is called
the systems state variable. If we know the value of systems state variable at any time t, then, with the value of
the input variable, we can calculate the value of every other
power variable in the system at that time t. We will first make
use of state variables to determine the initial value of an output variable in Sect.3.4. In Chap.7, we will introduce the
state-space which describes a dynamic system with multiple independent energy storage elements by a set of simultaneous first-order differential equations of the state variables,
rather than by a single higher-order differential equation.

3.2.3.9Derivation of a Differential System


Equation
We have extracted all information from the linear graph
which describes how the energetic elements are connected to
another. We have written element equations which describe
how the energetic elements behave individually. Finally, we
have written energy equations which can be used to establish the initial condition of any power variable in the system,
other than the input force which we control. By systematically extracting all of the information from the linear graph,
we can be assured that we have sufficient information to derive a differential system equation to describe the energetic
behavior of the system.
The verb derive connotes an expression of mathematic
logic. Once we have extracted and collected all of the mathematical statements of the energetic properties of a system,
we need effective algebra, more than mathematical logic. We
use the verb reduce to describe the process of elimination
by substitution, algebraic rearrangement, and occasional differentiation to arrive at a differential system equation. We
begin the reduction of this set of equations to an inputoutput
relationship, relating the applied force to our output of interest, by choosing any of the node, path, or element equations
except a trivial equation. Do not use the energy equations to
derive a differential system equation. The energy equations
are used to determine the initial conditions needed to solve
the system equation.
We will choose two output variables: the velocity of the
mass, v1g, and the force acting to accelerate the mass, FM.
Again, reduction of the energetic equations to a differential
system equation is achieved primarily through elimination
by substitution. Our objective is to eliminate all of the power
variables in the equation except for the input and the output variable. Once we have an equation written in terms of
the input variable, the output variable, time, and the element

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

128

Fig. 3.13a Translational mechanical mass-damper system


driven by a force source.
b Linear graph of the massdamper system driven by a
force source

x,v
F(t)

Mass
M

F(t)

Lubricating
fluid film
Damping b

Energetic Equations

Compatibility, Path Eq: v1g = v1g

Energy Eqs: E sys = E M

FM = M

dv1g

dt
1
E M = Mv12g
2

dt

F (t ) = M

F (t ) M dv1g
=
+ v1g
b
b dt

(3.26)

Check the units of the system equation in terms of the power


variables and time. The derivative operator and the element
and node subscripts do not affect units. They are removed to
simplify the notation:

[ F ] = M
M

dv1g

dt

[ F ] = bv
b

1g

[ F ] = M t

[ F ] = [bv ]

Ft
v

[ M ] =
F

[b] = v

F M dv1g
F M v
b = b dt + v1g b = b t + [ v ]

v F t v v
+ [v]
F
=
F

v F t

[v] = [v] + [v]

dv1g dFM
dF
dF (t ) d
dF (t )
= (bv1g ) + M
=b
+
dt
dt
dt
dt
dt
dt

F (t ) Fb FM = 0 F (t ) = Fb + FM
dv1g

F (t ) Fb FM = 0 F (t ) = Fb + FM F (t ) = bv1g + FM

Reduction i: Input F (t ), Output v1g

F (t ) = bv1g + M

F (t ) M dv1g
=
+ v1g
b
b dt


The units are in check. Unfortunately, this check can only


identify equations in which the units are inconsistent between terms. The system equation can still be erroneous, if
the units are consistent. However, it is definitely erroneous,
if the units are inconsistent.
Reduction ii: Input F (t ), Output FM

Continuity, Node Eq: F Fb FM = 0

Element Eqs: Fb = bv1g

g
parameters, we have completed the reduction. We will then
express the system equation in a time constant form, and
then check the units of the system equation in terms of the
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
graph drawn from it should be grouped with the set of energetic equations. The set of energetic equations is used first,
to formulate the differential system equation and, again, to
establish the initial conditions needed to solve the system
equation. The order the energetic equations are presented
is not important. It is important to organize the energetic
equations by type. If an equation is missing, it may not be
immediately apparent during a reduction. Be systematic in
extracting the information from the linear graph. Separate
the set of energetic equations from any subsequent algebra.
Combining statements of physical truth with possibly erroneous algebra is counter-productive. Finally, identify the
input variable and the power variable to be the output variable, that is, the objective of the reduction, at the top of the
reduction.

dv1g
dt

+ bv1g

dF (t )
dF (t ) b
dF
F dF
FM + M
= b M + M
=
M
dt
dt
dt
M
dt
M dF (t ) M dFM
=
+ FM
(3.27)
b dt
b dt

3.2 Introduction to the Linear Graph Method

129

Check the units:


Ft
v

[ M ] =

[b] = v

M dF M dFM
M F M F
b dt = b dt + [ FM ] b t = b t + [ F ]

F t v F F t v F
+ [F ]

=
v F t v F t

[F ] = [F ] + [F ]
The units are consistent.

3.2.4Examples Illustrating the Linear


Graph Method
We will illustrate the linear graph method by deriving system
equations for two more first-order translational mechanical systems, and a second-order translational mechanical
system. The systems contain elastic strain energy storage,
which is represented schematically as a spring. The familiar schematic symbol for a translational spring is shown in
Fig.3.14. Note that there are nodes at either end of a translational spring. The ends can displace independently of one
another. The deformation of the spring is the difference in
the displacements of the two ends. The rate of deformation is
the difference in the velocities of the two ends. Also note that
the magnitude of the spring force, FK, is the same at both the
ends. The spring force acts through the spring.
In an energetic model, a spring represents strain energy
storage. The strain energy may be stored in a machine component intended to act as spring but it need not be. Any machine loaded machine component will deform. It the deformation stores a significant amount of elastic strain energy,
then than energetic property of the physical system is represented schematically by a spring. A spring is identified by
its spring constant K. The machine design term spring rate
is not used in system dynamics because rate is a misnomer. The familiar expression relating force acting through a
spring to the deformation of the spring, Eq.1.7,
(1.7)
FK = K x12
is differentiated with respect to time, so as to express the relationship in terms of the power variables, force and velocity
of a translational mechanical system:
dF
= Kv
(1.42)
dt

FK

x 1 ,v1
Node 1

x 2 ,v 2

FK

Node 2

Fig. 3.14 A spring schematic symbol for translational damping

The product of the velocity difference across the spring and


spring constant K is the rate at which the force acting through
a spring changes.
The two example systems shown in Fig.3.15 differ in
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
mechanical element which is ideally rigid can neither store
strain energy nor dissipate energy. If the element is also ideally massless, it cannot store kinetic energy either. Consequently, a rigid, massless bar has no energetic properties.
Its purpose in an energetic model is to transmit force and
power. In the example system, the rigid and massless bar
connects the force source with the spring and damper. If the
physical system had the same configuration as the energetic
schematic, we would expect the bar to rotate during its transient response. However, an energetic schematic represents
the energetic properties of a physical system and need not
resemble the physical system. Both the strain energy storage
and the viscous damping in the schematic may, in fact, be attributes of a single machine component. Defining only translational displacement and velocity on the energetic schematic declares that we have restricted motion in the model
to transition. If there were rotation in the physical system
and a significant amount of energy was stored or dissipated
in rotation then we would add a rotational subsystem to the
model, using an interface element called a transducer, which
is introduced in Chap.6.
The system shown in Fig.3.15b is driven by a velocity
source. A translational velocity source is often a rod connected to the energetic system it is driving. It could be the
operating rod of a pneumatic or hydraulic piston under feedback velocity control. Another type of translational velocity
source is a linkage connected to another machine or machine
element which moves through a prescribed motion. Regardless of the details, the motion of the machine element acting
as a velocity source is either controlled or known. What is
unknown is the force, which must be applied by the machine
element to impose the specified velocity at its point of application. In short, we control or know the velocity of the
point of application. The dynamic response of the system
determines the force required to impose that velocity on the
system.
Velocity sources are unfamiliar to most students. They
are not used in dynamics because, in the general case, it is
impossible to apply a velocity source to a mass. A physi-

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

130
Fig. 3.15a Translational
mechanical spring-damper
system driven by a force source.
b Translational mechanical
spring-damper system driven
bya velocity source

x,v
K

F(t)

v(t)

x,v

Rigid and massless bar

cally real source can supply only a finite amount of power.


We shall see in Sect.3.4 that instantaneous application of
a step input would require an instantaneous change in the
velocity of the mass. A finite change in the velocity of the
mass means a finite change in the kinetic energy of the mass.
Power is the flow rate of energy. An instantaneous change in
the kinetic energy of a mass requires infinite power:

(1.28)
dt
0
Likewise, in the general case, a force source cannot act directly on a spring. We would need infinite power, if we were
to dictate the force acting through a spring by imposing a
step change on it, since the strain energy stored in a spring is
a function of the force acting through the spring. The argument can also be made using deformation. If a spring were to
deform a finite amount instantaneously, the end of the spring
would displace at infinite velocity. You may recall problem
statements from previous courses, in which a force was applied directly to a spring. This is impossible in the general
case, but not in two special cases. The first special case applies to a force from the source, acting through the spring
which is increased gradually over time, limiting the power
flow from the force source. We will use for this special case,
when we investigate the application of sinusoidally varying
forces to systems. The second special case occurs, when the
transient response is neglected, and only the steady-state is
considered. Steady-state occurs after the force source and the
spring have reached equilibrium. This special case is commonly used in machine design but is not used in system dynamics, since we are interested in the transient response of
systems.

3.2.4.1 Force Source-Spring-Damper System


We will now draw the linear graph of the force sourcespring-damper system, shown in Fig.3.15a. The first step
is to identify the nodes of the across variable. The across
variable in a translational mechanical system is velocity.

There are two velocity nodes associated with every energetic element. Except for a mass, the velocity nodes are at
either end of every element. Again, masses are special. The
second node associated with a mass is the inertial reference,
ground. Draw both nodes associated with each element on
the schematic, Fig.3.16a, then eliminate redundant nodes
of the same value of the across variable, Fig.3.16b. Rigid
objects have a single translational velocity, when motion is
restricted to one axis or dimension. The three nodes on the
rigid and massless bar are the same velocity and, therefore,
the same node.
Three ground nodes have been identified. The force
source, represented as a vector, must react against something. Since its reaction is not shown, we assume that it reacts against ground. Only one of two ground nodes on the
right end of spring K and damper b is needed. Note that
ground is also indicated schematically by the cross-hatching.
Draw the linear graph by first drawing and labeling nodes
1 and g. Then, add the elements by drawing lines (branches) between the respective nodes. Draw arrowheads on the
branches to define the positive direction of the through variable force for the element. Identify the branch with the parameter of the element, Fig.3.17.
The energetic equations of the system are categorized as
follows:
1. Continuity or Node Equations: Sum through variables
(forces) at nodes. Write one continuity equation less than
there are nodes. Omit the ground node.
2. Compatibility or Path Equations: Sum across variable
differences between two nodes along different paths.
Write no more compatibility equations than there are
holes in the linear graph.
3. Element Equations: Write one element equation relating
the two power variables of each passive element. There is
element equation for a source.
4. Energy Equations: Equate the energy stored in the system with the energy stored in the energy stored elements.
Write an energy equation for each energy storage element.

3.2 Introduction to the Linear Graph Method

131

Fig. 3.16a The nodes at both


ends of each element of the system are identified. b Redundant
nodes removed and the nodes
labeled

x,v
K
F(t)

K
F(t)

Fig. 3.17a Velocity nodes of the system shown in Fig.3.16. b Linear


graph of that system

Writing the set of energetic equations in a neat and orderly


manner saves time and reduces errors. Discipline yourself to
write the complete set of energetic equations, before beginning the reduction of the equations to a differential system
equation.
Energetic Equations

Rigid and massless bar

F (t ) Fb FK = 0 F (t ) = Fb + FK F (t ) = bv1g + FK
F (t ) =

b dFK
+ FK
K dt

Continuity or Node Eq: F (t ) Fb FK = 0


Compatibility or Path Eq: v1g = v1g
dFK
= Kv1g
dt

Energy Eqs: E sys = E K

F2
EK = K
2K

We will reduce the set of energetic equations to the differential system equations, which relate the input force to the
force acting through the spring, FK, and the force acting
through the damper, Fb. Recall that the process utilizes elimination by substitution, differentiating if necessary. The energy equations are not used in the reduction to a d ifferential

(3.29)

Check the units of the system equation in terms of the power


variables, F and v, and time t:
F
dFK
= Kv1g [ K ] =
dt
t v

F
Fb = bv1g [b ] =
v

b dFK
F (t ) =
+ [ FK ]
K dt
F t v F
+ [F ]
v F t

[F ] =

Element Eqs: Fb = bv1g

system equation. Start with a summation containing either


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

F(t)

Rigid and massless bar

x,v

b F
+ [F ]
t

[ F ] = K

[F ] = [F ] + [F ]

The units are consistent.


Reduction ii: Input F(t), Output Fb
F (t ) Fb FK = 0 F (t ) = Fb + FK
Differentiate both sides with respect to time, in order to substitute for FK:
dF (t )
dt

dFb dFK
+
dt
dt
dF (t )
dt

dF (t )
dt

dFb
+ Kv1g
dt

dFb
F
+K b
dt
b

This is the system equation, since it is written in terms of


the input and output variable, time, and the element param-

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

132

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


by clearing the coefficient from the undifferentiated, or zeroorder, output variable term:

x,v

v(t)

b dF (t ) b dFb
(3.30)
=
+ Fb
K dt
K dt
Check the consistency of the units in terms of the power variables and time:
F
dFK
= Kv1g [ K ] =
dt
t v

Fig. 3.18 Nodes of distinct values of the across variable velocity, identified on the energetic systematic

x,v

F
Fb = bv1g [b ] =
v

b dF (t ) b dFb
b F b F
+ [ Fb ]
=
+ [F ]

K t K t
K dt K dt
F t v F F t v F

=
+ [F ] [F ] = [F ] + [F ]
v F t v F t
The units check.

3.2.4.2Velocity Source Spring-Damper System


The linear graph of the velocity source spring-damper s ystem
shown in Fig.3.15b is drawn following the same procedure.
The first step is to identify the nodes of the across variable
translational velocity on the schematic of the energetic system. Draw nodes at either end of each element on the schematic, Fig.3.18. The velocity source is a power source. It
will exert the force necessary at its point of application to impose the specified velocity. Velocity sources must have a reaction, or they cannot develop the force necessary to impose
the specified velocity. The velocity source is represented as a
vector, indicating the nature of the source is either unknown
or unimportant. Place a ground node at the tail of the velocity
source vector. Next, eliminate redundant nodes of the same
value of the across variable velocity, if necessary. In this system, the only redundant nodes are ground nodes which we
are free to keep, Fig.3.19.
The linear graph is drawn by drawing and labeling the
nodes identified on the schematic, Fig.3.20a, and then adding branches which represent each energetic element between the nodes. The velocity source symbol is a circle in
the middle of the branch with a v in it, to identify the
sources power variable, which is known or controlled.
Draw arrowheads on the branches to define the positive
direction of the through variable force, Fig.3.20b. In the
spring and the damper, the across variable decreases in the
positive direction of the through variable. In this system,
the spring is compressing, when the velocity of node 1 is
greater than node 2. Likewise, the damper is in compression, when the velocity of node 2 is positive and, hence,
greater than ground velocity of zero. Sources are special

v(t)
g

K
1

Fig. 3.19 Node labeled. A ground node is needed for the velocity
source to react against

since they supply power to the system. The across variable


increases in the positive direction of the through variable in
a source element.
Write the complete set of energetic equations before beginning the reduction to a system equation. Write one continuity or node equation less than there are nodes. There are
three distinct velocity nodes in this system, since we drew a
redundant ground node for convenience. Therefore, we will
write two node equations. We will omit the continuity equation for the ground node. There is force acting through the
velocity source. It is an unknown, which is established by the
response of the system.
Energetic Equations
Continuity or Node Eqs: Fsource = FK

FK = Fb

Compatibility or Path Eqs: v1g v (t ) = v12 + v2 g


Element Eqs: Fb = bv2 g
Energy Eqs: E sys = E K

dFK
= Kv12
dt

EK =

FK2
2K

We will reduce this set of energetic equations to the differential system equations, which relate the input force to the
force acting through the spring, FK, and the velocity drop
across the spring, v12. Again, start with a summation containing either the input or the output variable, if possible. Make
successive substitutions to eliminate any variable, other than
the input and output variables and time. Element parameters
are not variables, and they are not eliminated.

3.2 Introduction to the Linear Graph Method


Fig. 3.20a The nodes identified on the energetic schematic,
Figs.3.18 and 3.19 are drawn
and labeled. b Branches representing the energetic elements
are drawn between the respective
nodes, labeled with the element
parameter, and oriented with
an arrowhead pointing in the
positive direction of the through
variable, force

133

v(t)

g
Reduction i: Input v(t), Output FK
v (t ) = v12 + v2 g v (t ) =

1 dFK Fb
1 dFK FK
+
v (t ) =
+
K dt
b
K dt
b

This is the system equation. Clear the coefficient from the


zero-order (constant) term to express the system equation in
a time constant form:
b dFK
+ FK
bv (t ) =
(3.31)
K dt
Check the consistency of the units in terms of the power variables and time:
dFK
= Kv12
dt

[K ] = t v

F
Fb = bv2 g [b ] =
v

b dFK
bv (t ) = K dt + [ FK ]

b F
+ [F ]
t

[bv ] = K

F F t v F
+ [F ]
v=
v v F t

[F ] = [F ] + [F ]

The units are consistent.


Reduction ii: Input v(t), Output v12
v (t ) = v12 + v2 g v (t ) = v12 +

Fb
F
v (t ) = v12 + K
b
b

Differentiate with respect to time to eliminate FK, using the


element equation for the spring:
dv (t ) dv12 1 dFK
dv (t ) dv12 1
=
+

=
+ Kv12
dt
dt b dt
dt
dt b
This is the system equation. Express it in a time constant
form:
b dv (t ) b dv12
=
+ v12
(3.32)
K dt
K dt

g
Check the consistency of the units in terms of the power variables and time:
dFK
= Kv12
dt

[K ] = t v

Fb = bv2 g

b dv (t ) b dv12

=
+ [ v12 ]
K dt K dt

[b] = v

b v b v
K t = K t + [ v ]

F t v v F t v v

=
+ [v]
v F t v F t

[v ] = [v ] + [v ]

The units check.

3.2.4.3 Force Source-Spring-Mass-Damper System


Two equivalent energetic schematics of a spring-mass-damper system driven by a force source are shown in Fig.3.21.
The schematic in Fig.3.21a represents the damping b of the
system in the shear of a viscous lubricating film. The schematic shown in Fig.3.21b shows mass M supported on ideal
frictionless rollers. The systems damping b is represented
by a dashpot. Both of these schematics yield the same linear graph, because they have the same elements connected
between the same nodes of distinct values of the across variable velocity, as we will demonstrate. The schematics can be
used interchangeably.
The first step in drawing a linear graph from an energetic schematic is to find and label nodes of distinct values
of the across variable, translational velocity in this system,
Fig.3.22. Sources, springs, and dampers have velocity nodes
at either end of their schematic symbol. In the case of a force
source represented as a vector, place a ground node on the
vectors tail. Force sources must have a reaction, or they cannot apply a force. Masses are special. Since their only property is storage of kinetic energy, they must be rigid. A mass
has a single velocity node on the element. A mass other velocity node is its inertial reference, ground.
Write the complete set of energetic equations, before beginning the reduction to a system equation. Write one continuity or node equation less than there are nodes. Therefore,

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

134
Fig. 3.21a Schematic of a
force source spring-mass-damper
system, with the damping
represented as a lubricating film
supporting the mass. b Schematic
of a force source spring-massdamper system, with the damping
represented by a dashpot

x,v

F(t)

F(t)

Frictionless
rollers

Fig. 3.22a, b Energetic schematics of a force source springmass-damper system with nodes
at either end of the force source,
spring, and damper

M
K

Lubricating fluid
Damping b

x,v

x,v

F(t)

F(t)

g g

we will write the node equation for node 1, and omit the
ground node. There are three holes in the linear graph. If
the linear graph had two elements in each path from node 1
to ground, then three useful compatibility or path equations
could be written. Since there is only one element in each
path from node 1 to ground, only one unique equation can be
written, and it is a trivial equation.
Energetic Equations
Continuity or Node Eq: F (t ) = Fb + FK + FM
Compatibility or Path Eq: v1g = v1g
dFK
= Kv1g
dt

Energy Eqs: E sys = E K + E M

EK =

FM = M

FK2
2K

EM =

dv1g
dt
1
Mv12g
2

We will reduce the energetic equations to system equations


for the two energy storage variables, FK and v1g. These are
the systems state variables, because they, along with the
input variable, determine the energetic state of the system.
Reduction i: Input F(t), Output FK
We start with continuity equation, because it is a summation that includes the input and output variables. Do not

1 1

g
Lubricating fluid
Damping b

Element Eqs: Fb = bv1g

x,v

Frictionless
rollers

substitute for either the input or output variable. Substitute to


eliminate all other power variables:
F (t ) = Fb + FK + FM F (t ) = bv1g + FK + M
F (t ) =

dv1g
dt

d 1 dFK
b dFK
+ FK + M

dt K dt
K dt

F (t ) =

M d 2 FK
b dFK
+ FK +
K dt 2
K dt

F (t ) =

b dFK
M d 2 FK
+ FK +
K dt
K dt 2

This is the system equation. Express it in standard form by


rearranging the derivatives into decreasing order, and clearing the coefficient from the highest-order derivative:
F (t ) =

M d 2 FK b dFK
+
+ FK
K dt 2
K dt

K
K M d 2 FK K b dFK K
F (t ) =
+
+
FK
M
M K dt 2
M K dt
M
d 2 FK
K
b dFK K
F (t ) =
+
+
FK
(3.33)
2
M
M dt
M
dt

3.2 Introduction to the Linear Graph Method

135

Check that the units of all terms are consistent. Substitute


equivalent units for the element parameters:
Fb = bv1g

[b] = v

FM = M

dv1g
dt

dFK
= Kv1g
dt

[K ] = t v

Ft
v

[ M ] =

2
1 dF (t ) d v1g b dv1g K

2 +
+ v1g
M
dt
M
dt
M
dt

1 F v b v K
M t = t 2 + M t + M
v F v F v

= 2+
F t t t v F t

2
K
d FK b dFK K

M F (t ) = dt 2 + M dt + M FK

The units check.

F v
F F v F F v
F = 2 +
F

+
t
t
t
v
F
v
F
t
v
F
t
t
t

3.2.5Summary of the Introduction to Linear


Graphs

F F F F
t 2 = t 2 + t 2 + t 2
The units check.
Reduction ii: Input F(t), Output v1g
We again start with a continuity equation, because it is a
summation which includes the input variable:
dv1g
dt

dv1g dFK d dv1g


dF (t )
=b
+
+ M
dt
dt
dt
dt
dt
dv1g
d 2 v1g
dF (t )
=b
+ Kv1g + M
dt
dt
dt 2
This is the system equation. Express the system equation in
standard form by rearranging the derivative terms, in order
of decreasing order of differentiation, and clearing the coefficient of the highest-order output variable term:
d 2 v1g
dv1g
dF (t )
=M
+b
+ Kv1g
2
dt
dt
dt
2

1 dF (t ) d v1g b dv1g K
=
+
+ v1g
(3.34)
M dt
M dt
M
dt 2
Check the units.
Fb = bv1g

dFK
= Kv1g
dt
F

[b] = v

FM = M

dv1g
dt

Ft
v

[ M ] =

v F v
v
+
t t v F t

v v v v
t 2 = t 2 + t 2 + t 2

K F b F K
M F = t 2 + M t + M F

F (t ) = Fb + FK + FM F (t ) = bv1g + FK + M

[K ] = t v

The most important step in developing a differential system


equation is modeling the energetic system. The modeling
process requires judgment, in order to retain the significant
energetic attributes of a system, while neglecting those which
can be considered insignificant. Understanding dynamic systems is a prerequisite for modeling dynamics systems. Towards that end, energetic models are provided in this text as
schematics with the energetic attributes indicated.
Begin drawing a linear graph from the schematic of an
energetic model by identifying the nodes of distinct values of
the across variable. Mark and label them on the schematic to
prevent the omission of any nodes. An error in this step carries forward through the subsequent effort. In translational
mechanical systems, the across variable is translational velocity. There must be an across variable difference associated
with every element.
The depiction of the source, acting on a system as a vector, is a modeling convention, which indicates that details
of how the power is generated and applied to the system
are either unimportant or unknown. The input forces point
of action has a distinct velocity. The tail of the force vector
must also have a distinct velocity, which is different from
the head. Mechanical sources must react against something, in order to apply a force to a system. If no element is
shown for the reaction, then assume the source is reacting
against ground. A mass is rigid and, thus, can have only
one velocity. The velocity difference across the fluid film is
the velocity difference between the mass and ground. The
velocity difference in a dashpot is between the piston and
the cylinder.
Draw the nodes and label them. Then draw the branches, which represent the energetic elements between the
nodes identified on the schematic. Energy storage and dissipation elements are represented by a line with an arrowhead in the middle, to define the positive direction of the

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

136
Fig. 3.23a, b Energetic schematics of a force source springmass-damper system redundant
nodes on the ideally rigid mass
removed

F(t)

F(t)
g g

g
Lubricating fluid
Damping b

x,v

F(t)

x,v

Frictionless
rollers

We will continue to develop the linear graph method in


Chaps.4, 5, and 6, as we introduce additional energetic elements.

3.3 The Heaviside Unit Step Function


The Heaviside unit step function, us (t ta ), has one of two
values, Fig.3.25:

Fig. 3.24a Nodes of distinct values of the across variable velocity of


the force source spring-mass-damper schematics of Fig.3.23. b Linear
graph of those schematics

through variable, and identified by the elements parameter. Force is the through variable in a translational mechanical system. Sources supply power to a system. Their
symbol is a branch with a circle. The variable within the
circle identifies the type of source. Only one of the two
power variables can be controlled by a source. The other
power variable is determined by the power drawn from
the source by the system. A unique aspect of a source is
that across variable increases in the positive direction of
the through variable. Conversely, in energy storage and
dissipation elements, the across variable decreases in the
positive direction of the through variable.
A linear graph is an energetic network. Continuity, or
node, equations and compatibility, or path, equations describe the configuration of the network. Node equations sum
through variables at nodes. One less node equation is written
than there are nodes. We will omit the node equation for the
ground node. Path equations that equate the sum of across
variable drops between two nodes along different paths. The
element, or constitutive, equations relate the two power variables of an element to each other. The element equations are
written for the assumed positive direction defined by the arrowhead on the elements branch. Equations for the amount
of energy in an element and for the amount of energy in the
system are used to establish the initial conditions needed to
solve the differential system equation, not to derive the system equation.

We choose what time to define as t = 0. Consequently, we


define time, t = 0, to be the beginning of the time period of
interest to us. We may define t = 0 to be the time at which
a step function transitions (or turns on). In order to turn
on a step input at an arbitrary time ta other than t = 0, we
subtract the constant ta from the independent variable t. If
ta is positive, the Heaviside step functions arguments transition from negative to positive is delayed, or shifted right
on the time axis, Fig.3.26. Subtracting a positive constant
ta from time variable t keeps the argument of the unit step
function, (t ta ), negative until time, t = ta . Conversely, if ta
is negative, subtracting the negative constant from the time
variable, or, equivalently, adding a positive constant, shifts
the transition of the step function to that negative time ta.
Graphically, adding a positive constant to the time variable
causes the Heaviside step functions argument to become
positive earlier, shifting the transition of the step function
left on the time axis.
The Heaviside step function is an idealization used to
model step inputs. An event which is truly described by a
Heaviside step function cannot actually happen in physical
systems, except on the scale of subatomic particle, because
the transition of the Heaviside step functions value from
zero to one is instantaneous, Fig.3.27. On the macroscopic
scale of everyday life, physical changes can be extremely
rapid, but they cannot be instantaneous. The power variables
in our models can change instantaneously, except for the

us (t ta ) = 0 for t < ta
us (t ta ) = 1 for t > ta

(3.35)

3.3 The Heaviside Unit Step Function

137

us(t)
1

us(t)

u s(t) = 1

u s(t) = 0

time

t = 0-

Fig. 3.25 The Heaviside unit step function. The unit step function
transitions from 0 to 1, when its argument transitions from negative
to positive

t=0

t = 0+

time

Fig. 3.27 Ideal Heaviside unit step function

us(t-t a )

us(t-t a ) = 1

u(t)
1

u(t) = 1

us(t-t a ) = 0

t-t a < 0

ta

t-t a > 0

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

state or energy storage variables, which require time. The


value of a state (energy storage) variable cannot change instantaneously without infinite power, since there would be a
finite change in the energy in the storage element over zero
time. Energetic models are simplified representations of real
physical systems. An actual physical machine component
used to apply an input to a system possesses all of the energetic attributes, energy dissipation and both modes of energy, to some extent. We have neglected all but the most significant attributes to create a simplified model. In practice,
we cannot generate ideal step inputs because of the energy
storage attributes of our actual power sources. The closer we
examine power sources, the clearer it is that they are, in fact,
dynamics systems.
Fortunately, if the input variable changes from zero to
a constant value over a time which is small relative to the
timescale of the transient response of the system it is driving, then the input looks like a step change to the system.
There are many circumstances, when an input variable can
be reasonably modeled as an instantaneous step change. A
Heaviside step function can be used to model an input variables transition, when the duration of the transition of an
actual physical input power variable t is brief, relative
to the timescale of the systems response measured, by either its time constant, , or period of oscillation, T. When
duration t of the transition of a physical input variable is
short relative to the systems timescale, then the proportion

u(t) = 0
t<0

t < t

time

Fig. 3.28 Realistic unit step transition

of the transient response, which occurs during transition of


the input is small. When almost all of the systems transient
response occurs after the actual transition of the input variable, the Heaviside step function is an excellent model for
the application of the input. On the other hand, when the
duration t of the transition is a substantial fraction of the
systems timescale, either the time constant, , or the period
of oscillation, T, then the system will have begun to respond
to the input, before the transition is complete, and modeling
the input as a Heaviside step function would be inaccurate.
The input model would need a transitional period for the systems response to be accurately predicted, Fig.3.28.

3.3.1Differentiation of the Heaviside


Unit Step Function
Many of our differential system equations have differentiation of the input function. Consequently, we must address the question of how to handle differentiation of the
Heaviside step function. The derivative of the Heaviside
unit step function with respect to time is easy to evaluate
for t ta , since the step functions value is a constant, either zero or one:


dus (t ta )
dt

= 0 for t ta

(3.36)

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

138

(t)

(t) = 0
t<0

t = 0-

t=0

(t) = 0
t>0
+
t=0

time

Fig. 3.29 Unit impulse or Dirac delta function, (t)

The instantaneous transition of the Heaviside step function is a discontinuity at which its derivative with respect to
time becomes infinite. Mathematically, the derivative of the
unit step function is the unit impulse function, also known
as the Dirac delta function, (t), which has infinite amplitude and zero width (duration), Fig.3.29. Paul Dirac was a
British physicist influential in the development of quantum
mechanics.


(t ) = 0 for t 0

(t ) = for t = 1

(3.37)

The unit impulse function is a truly bizarre concept invented


for mathematical convenience. The term unit refers to the
fact that the value of the integral of the unit impulse equals one:


(t ) dt = 1

(3.38)

By itself, this makes sense, since the unit impulse is the derivative of the unit step and integration is the inverse operation of differentiation.
Infinity as the value of the derivative of a function which
transitions instantaneously also makes sense mathematically.
It is a singularity, where we are dividing finite change by
zero time:
dus (t )
( 0) =
(3.39)
dt t = 0
Although we can calculate a value of infinity, we cannot
physically create the time derivative of a power variable with
infinite magnitude. The bizarre aspect of a unit impulse is
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 Heaviside unit step function. What is physical effect on a system
of an input term which differentiates the Heaviside step
function? Fortunately, the physical effect is straightforward.
There is none. The input to the actual physical system is a

Heaviside unit step. The differential system equation for a


power variable within the system has the derivative of the
Heaviside step as an input term, dus (t )/ dt. The initial conditions of the system are determined by the energy in the system prior to the application of the Heaviside unit step functions transition and the value of the input the instant after it
has transitioned, as we will see in Sect.3.4. Recall that the
input terms were only used in the method of undetermined
coefficients to establish the particular solution, which is the
steady-state solution. The roll of the derivative of the any
input term is to establish its contribution to steady-state response (the particular solution). As a result, we ignore its derivative at time t = 0. At first, this statement may be surprising. How could an infinite value of an input term not have an
immediate effect on the response of the output variable? An
input term to a system equation which is the derivative of the
Heaviside step function does not hammer a system with an
impulse of infinite magnitude and zero duration. The input to
the system is a step of finite magnitude.
We shall see in Chap.8 that the value of the derivative of
Heavisides step function at the instant of transition, when its
argument equals zero at t = ta , is problematic in numerical
solutions. Heavisides original definition of the step function, Eq.3.35, has no value at the instant of transition. In
other words, the function and its derivative were undefined at
time t = ta . Heavisides original definition of the step function contains two different types of discontinuities. Heavisides step function has a jump discontinuity, which means
there is one value as t = ta is approached from positive time,
and a different value as t = ta is approached from negative
time. Heavisides definition also has a point discontinuity,
meaning the function is missing a value at a point. Specifically, the function has no value at time t = ta . A number of
modern definitions of Heavisides step function which have
added a value at time t = ta or values in the vicinity of t = ta
to eliminate the point and jump discontinuities. If you have
worked with Mathcad, you may have noticed that Mathcads
Heaviside step function adds two points in the transition interval. Eliminating the point discontinuity of the step function is important for machine computation, since numerical
computations need functions to return values.
Although we will not do so, a finite pulse can be expressed
as a continuum of unit impulses in the time-domain. We will
have use for the unit impulse as an input in the Laplace-
domain, as we shall see in Chap.10. It serves to initiate the
homogeneous response of a transfer function.

3.4 Initial Conditions


In a textbook problem, the initial conditions needed to solve
a differential equation are given. In system dynamics, the
energetic state of the system at the instant before the input

3.4 Initial Conditions

139

Fig. 3.13a Translational mechanical mass-damper system


driven by a force source. b Linear
graph of the mass-damper system
driven by a force source

x,v
F(t)

Mass
M
1

Lubricating
fluid film
Damping b

F(t)

F0 us(t)

t, sec

Fig. 3.30 Step input driving the mass-damper system of Fig.3.13 and
the spring-mass-damper system of Fig.3.31

is applied, and the values of input are used to determine the


initial conditions needed to solve the system equation.

3.4.1 Initial Condition, First-Order System


We shall use the mass-damper system of Fig.3.13 to illustrate
how to determine the initial conditions needed to solve the
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.
Figure3.13, the energetic equations, and the system equations
are reproduced below from Sect.3.2.3.9 for reference.
Energetic Equations
Continuity or Node Eq: F Fb FM = 0
Compatibility or Path Eq: v1g = v1g
Element Eqs: Fb = bv1g

FM = M

Energy Eqs: E sys = E M

EM =

dv1g
dt

1
Mv12g
2

System Equations: Eqs.3.26 and 3.27


F (t ) M dv1g
=
+ v1g
b
b dt

M dF (t ) M dFM
=
+ FM
b dt
b dt

F(t)
N
F0

Let us say that the system is de-energized before the step


input, F (t ) = F0 us (t ), Fig.3.30, is applied.
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.
The order of a differential system equation equals the number of independent energy storage elements in the system.
The mass-damper system has a single energy storage element, the mass. It is a first-order system. The initial condition needed is the value of the output variable at the instant
after the input is applied at time t = 0+ .
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
the system, then the velocity of the mass is zero. Expressing
these statements mathematically,

E sys ( 0 ) = E M ( 0 )
and

E M (0 ) =

1
Mv12g 0
2

( )

( )

v1g 0 = 0

The energy storage variable, v1g, is the state variable of the


system. State variables have two important attributes. First, if
we know the value of the state variable at any time t, then we
can calculate any other power variable in the system at that
time by using its value and the value of the input value, which
is always known. Second, state variables are the only variables
in a system which cannot change instantaneously. State variables represent the amount of energy stored in energy storage
elements. Energy flows at a finite rate, which is power. A finite
change in the amount of energy stored in an element requires a
finite amount of time for the energy flow to occur. An instantaneous change in the amount of energy in an element would
require infinite power, which is physically impossible:
E E

P
(1.40)
dt
0

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

140

Since a state variable cannot change instantaneously, if we


know the value of the state variable the instant before a step
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
t = 0+. In this example, the system is de-energized at the instant before the step input of force, F(t), is applied. The force
source can only provide finite power. The energy transferred
to the system is zero over the infinitesimal time between
t = 0 and t = 0+ when the step transition is on. The mass
does not change energy. Consequently, the state variable, the
velocity of the mass, v1g, does not change between t = 0
and t = 0+ either:

E sys ( 0 ) = 0 = E sys ( 0+ ) E M ( 0 ) = 0 = E M ( 0+ )
1
1
Mv12g 0 = 0 = Mv12g 0+
2
2

( )

( )

( )

( )

v1g 0 = 0 = v1g 0+

As it happens, the state variable, v1g, is an output variable of


interest, for which we have established an initial value:

( )

(3.41)
v1g 0+ = 0
Regardless of whether or not the state variable is the output
variable, the first step in calculating the initial condition of
any output variable is always to determine the value of the
state variable at time t = 0 , the instant before the input is applied. 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
(state) variable, v1g (0+ ) = 0, and knowing the value of the
input at time t = 0+ ,

( )

( )

(3.42)
F 0+ = F0 us 0+ = F0
We can now establish the initial value of every other power
variable in the system by using algebraic energetic equations. We need to establish the value FM (0+ ).
We cannot use the differential elemental equation for the
mass, because we do not know the value of the derivative
of the velocity of the mass at time t = 0+ . We can calculate
the derivative, but we do not know it. Avoid the sophomoric
mistake of confusing knowledge of the value of a variable at
an instant in time with the variable being constant and concluding that the derivative of the variable is zero. We know
the value of the velocity of the mass at time t = 0+ is zero but
will not stay zero.
Perusing the energetic equations for FM, we see that the
only other appearance of FM is in the continuity or node
equation:
F Fb FM = 0

Evaluating this equation at time t = 0+ and rearranging the


expression yields

( )

( )

( )

F 0+ Fb 0+ = FM 0+

We know F (0+ ) = F0 , but Fb (0+ ) is unknown. We will


eliminate it by substitution. Reviewing the energetic equations, we see our only choice is the use of the elemental
equation for the damping Fb = bv1g, expressed in terms of
the known value of the state variable,

( )

( )

( )

F 0+ bv1g 0+ = FM 0+
Substituting for F (0+ ) and v1g (0+ ) ,


( )

F0 b v1g 0+

( )

= FM 0+

( )

FM 0+ = F0

(3.43)

we have the initial value of the force acting to accelerate


the mass. Our result makes sense. If the mass is motionless,
then the fluid film is not in shear, and there is no damping
force acting on the mass. All of the applied force acts to
accelerate the mass at the instant after the step input is applied. The division of the applied force, F(t), between the
force acting to accelerate mass FM and damping force Fb is
dynamic. The division of F(t) between Fb and FM changes
with time.

3.4.2 Initial Conditions: Second-Order System


We shall use the force source-spring-mass-damper system in
Fig. 3.31 as a second-order system to determine the initial
conditions needed to solve the system equations for the force
acting through spring FK, Eq.3.33, and the velocity of mass
v1g, Eq.3.34. The energetic schematic in Fig.3.23a, its linear
graph in Fig.3.24b, and energetic equations are reproduced
below for convenience. The system is driven by the step
input F (t ) = F0 us (t ) , shown in Fig.3.30.
Energetic Equations
Continuity or Node Eq: F (t ) = Fb + FK + FM
Compatibility or Path Eq: v1g = v1g
Element Eqs: Fb = bv1g

dFK
= Kv1g
dt

Energy Eqs: E sys = E K + E M

EK =

FM = M
FK2
2K

EM =

dv1g
dt
1
Mv12g
2

3.4 Initial Conditions

141

Fig. 3.31a Force source


spring-mass-damper system of
Fig.3.23. b Linear graph of the
force source spring-mass-damper
system

x,v

F(t)

F(t)
g

d FK
K
b dFK K
F (t ) =
FK
+
+
2
M
M dt
M
dt

(3.33)

We need the value of the output variable, FK (0+ ) , and its


derivative, dFK (0+ )/dt , the instant after step input is applied to the system at time t = 0+ . In order to derive the
initial conditions, we must know the amount of energy
in the system, and the location of that energy the instant
before step input is applied to the system at time t = 0 .
We will assume that the system is de-energized before the
step input is applied, E sys (0 ) = 0. Since E sys = E K + E M ,
and strain energy and kinetic energy must be positive due
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
have established the values of the energy storage, or state
variables, at time t = 0 :

E K (0 ) =

( )=0

FK2 0
2K

E M (0

1
= Mv12g 0 = 0 v1g 0 = 0
2

( )

( )

v1g (0 ) = 0 and v1g (0 ) = v1g (0+ ) v1g (0+ ) = 0


The only other variable in a system we know at time
t = 0+ is the value of the input, since we control it,
F (0+ ) = F0 us (0+ ) = F0 .
The initial conditions are determined by using these
known values, the energetic equations, and algebra. The output variable of system equation, FK, is a state (energy storage) variable. We assumed the system to be de-energized at
time t = 0 and have established the values of the state variables to be zero then and at the next instant, time t = 0+ . We
have established one of the two initial conditions we need,
namely,

( )

FK 0+ = 0
We now peruse the energetic equations and see that dFK /dt is
a term of the springs element equation:

( )

( )

It is impossible for the amount of energy stored to change


instantaneously during the transition of the Heaviside step
input, since any finite change over infinitesimal time would
require infinite power. The energy storage, or state, variables
are the only variables in a dynamic system which cannot
change instantaneously. If we know their values prior to the
application of a step input to a system, we know their values
immediately after the application of the step, because they
cannot change instantaneously:

( )

Likewise,

FK 0 = 0

and

( )

FK 0 = 0 and FK 0 = FK 0+

( )

FK 0+ = 0

g
Lubricating fluid
Damping b
What are the initial conditions needed to solve the system
equation, Eq.3.33?

dFK
= Kv1g
dt
The velocity of mass v1g is a state (energy storage) variable.
We have established at time t = 0+ , v1g (0+ ) = 0. Substituting,
dFK (0+ )
= Kv1g (0+ )
dt
We will now find the initial conditions needed to solve the
system equation, Eq.3.34, again assuming that the system is
de-energized prior to the application of the step input.


2
1 dF (t ) d v1g b dv1g K
=
+
+ v1g
M dt
M dt
M
dt 2

(3.34)

We have already established the values of the state (energy


storage) variables, FK (0+ ) = 0 and v1g (0+ ) = 0 . The system

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

142

equations output variable is v1g, so we have one of the two


initial conditions we need:

( )

v1g 0+ = 0
The second initial condition needed is the value of
dv1g (0+ )/dt . Checking the energetic equations, we find it in
the element equation for the mass:
FM = M

dv1g

dt

( )=

dv1g 0+

1
FM 0+
M

( )

dt

We must now establish FM (0+ ) in terms of the known


values of input, F (0+ ) = F0 , and the state variables,
FK (0+ ) = 0 and v1g (0+ ) = 0 . Begin by substituting for FM:

( )=

dv1g 0+
dt

( )=

dv1g 0+
dt

1
FM 0+
M

( ( )

( )

( )=

dt

( ))

1
F 0+ Fb 0+ FK 0+
M

( )=

dv1g 0+
dt

( ( )

( )

( ))

1
F 0+ bv1g 0+ FK 0+
M

( )

1
b
F 0+
v1g 0+
M
M

( )

( )=

dv1g 0+
dt

( )

If the system is of first order, then the characteristic equation


will be a first-order polynomial
s + a = 0 s = a
yielding the homogeneous solution
(3.45)
X h (t ) = Ae at

Eliminate Fb. Checking the energetic equations, we find that


the dampers element equation is written in terms of the state
variable, v1g, Fb = bv1g :
dv1g 0+

the input and output variables (the system equation) is a firstorder differential equation. If the equations which describe
the elements of the system are linear equations, then the system equation will be a linear differential equation, and can be
solved by determining the unknown coefficients of a general
solution. The homogenous solution is always an exponential
function of the form

X h (t ) = Ae st
(3.44)

FK 0+
M

( )

1
F0
M

Our result tells us that at the instant after the step input is
applied, time t = 0+ , all of the applied force is used to accelerate the mass. This makes sense, because no force can be
transferred to the spring or the damper, until the mass moves.

where the exponent is known and is a function of the parameters of the dynamic system. Coefficient A is unknown and
dependent on both the initial conditions and the input to the
system.
It is convenient to express the homogeneous response in
terms of the inherent or characteristic timescale of the firstorder system, known as the systems time constant, The
time constant, , of a first-order system equals the absolute
value of the inverse of its eigenvalue, , which, again, is its
characteristic value, s
t
1
1
(3.46)
X h (t ) = e st = e where = =
s

The particular solution has the same functional form as the


input driving or forcing the system. When the input is the
product of a constant and the Heaviside step function, which
is a constant after its transitions from zero to one,
Input (t ) = F0 us (t )
then the particular solution is a constant,

3.5 First-Order Step Responses

(3.47)
X p (t ) = C

The response of a system to a step input reveals the underlying nature of the dynamic system. Dynamic systems with
one independent energy storage element can only have
two basic responses to a step input. The system can either
monotonically gain energy from the source or lose energy
to the source. There can be no oscillations. Oscillatory step
responses are caused by energy flow between energy storage
elements within the system.
Systems with only one independent energy storage element are called first-order, because the equation relating

Physically, the particular solution is the response of the system after the effects of the initial conditions and the homogeneous response have decayed to zero. In the ideal case, time
must approach infinity for the effects of the initial conditions
to decay to zero. In practice, the steady-state response of a
system is defined as when there is no more discernible effect
of the initial conditions and all that remains is the response to
the input. In the particular case of a response of a system to a
step input, the system is declared to be in steady-state when
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
system

Decay

__
-t

__
-t

1- e

2 3 4 5
time, Time Constants

1+C

2 3 4 5
time, Time Constants

1+C

__
-t

__
-t

Offset Decay

e + C
C
0

Growth

(1- e ) +C
C
0

2 3 4 5
time, Time Constants

power variables in the system. The steady-state response of


all power variables in any system to a step input is to reach
a constant value. In practice, steady-state is defined to be
reached in time equal to a finite multiple of the time constant
of the system, t = 4, 5, or 6. The multiples of used depends on the linearity of the system, the quality of the signal,
and the precision of the measurement.
There is no further change with time in a system when
it reaches steady-state in response to a step input. This is
both an observed physical truth and a mathematical fact.
The particular solution, which is the steady-state response
of the system, has the same functional form as the input.
The Heaviside step input remains constant after its transition. Hence, the particular solution or steady-state response
of a linear system to a step input is a constant. Any power
variable in the system may be chosen as the output variable. All power variables in the system will be constant in
steady state.
The general solution to the differential equation describing the dynamic response of a first-order system is constructed by adding the homogeneous and particular solutions,
X (t ) = X h (t ) + X p (t ) = Ae at + C
(3.48)
There is no need to use this formal solution method to determine the response of a first-order system to a step input. Restricting the system to the first order and the input to a Heaviside step function limits the possible responses of a systems
output variable. The step response of all power variables in
a first-order system is either a stable exponential growth or
decay. In a given first-order system, some variables grow
while others decay. Importantly, all changes occur at the same
rate and at the same time. All power variables in the system
change simultaneously. The step responses of all variables in

Offset Growth

2 3 4 5
time, Time Constants

the system have the same time constant, . There is only one
time constant in a first-order system. However, responses of
individual power variables in a system can be in opposite directions. That is, a variable may grow, while another variable
may decay, as the system reaches its new equilibrium.
All first-order step responses are based on the same ext

ponential function, e , where e is Eulers number,


e = 2.7183 2.72, the base of the natural logarithm, and is
the time constant. The time constant must have units of time
so that the exponent is dimensionless. The time constant, ,
provides the timescale of the response. Since time in the exponent is divided by time constant , slow systems have large
time constants. Small time constants lead to fast responses.
A fast system is not necessarily a system which reaches a
high velocity. A fast system reaches steady-state quickly, and
a slow system reaches steady-state slowly.
There are only four solutions for the step response of a
first-order system, Fig.3.32: two growth equations and two
decay equations. The output variable can grow to a constant
value from either zero or a non-zero initial value C. The output variable can either decay from an initial value to zero
or to a non-zero constant value C. The constant C is a vertical shift or offset added to the exponential decay or growth
function.
One formulation of the step response of a first-order
system is


( ( )

X (t ) = X 0+ X ( ss ) e + X ( ss )

(3.49)

where X (0+ ) is the value of the output variable at the instant


after the step input is applied and X ( ss ) is the steady-state
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.
It is not a functional relationship in the sense that it does
not describe the response of a system. It is impossible to
visualize the shape of this function until X (0+ ) and X ( ss )
are known. In other words, because Eq.3.49 describes all
possible solutions, it has no specific meaning until the values have been determined. Second, and more importantly,
reliance on this formula obscures the essential point that
the change of any variable within a first-order dynamic systems is caused by the flow of energy into or out of the system. 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
and, hence, in the kinetic energy of the wheel. We can hear
the change in speed. Hence, we can hear the wheel gaining energy. When the wheel gains energy, we know that we
can describe the step response of its energy storage variable,
the wheels angular velocity, by a growth function. When it
loses energy, we can describe the wheels angular velocity
by a decay function. If the wheel is rotating in the negative
direction and gains energy, it makes more physical sense to
describe its step response as a growth to a negative steadystate value than as decay attributed to a negative steadystate value.
Understanding dynamic responses of physical systems
requires moving beyond a purely formulaic approach. Developing the ability to visualize response functions of a system to step input allows one to make productive use of ones
understanding of dynamic system behavior, even when there
is insufficient information to formulate a mathematical response.

3.5.1Categorization of First-Order Step


Responses
Scaling and normalizing are inverse mathematical operations. To formulate a response function, we multiply an exponential decay or stable growth function, which has a range
of zero to one, by a constant factor which scales the output variable and gives it physical units. The constant is the
dynamic range of the output variable. Normalization is accomplished by dividing observed data by the same constant
factor to reduce its maximum value to unity and remove the
physical units. The scaling and normalizing by multiplying
or dividing by the dynamic range affects only the ordinate
(vertical axis). The abscissa (time axis) is normalized the
time constant, , as shown in Eqs.3.50 and 3.51, and then
by measuring or expressing time in units of time constants.
Division of the time variable in the exponent by the time
constant is not an additional step. It is an essential
part of the
t
homogeneous solution, vh (t ) = Ae st = Ae . The exponent

V(0+)
63.2%

V(t)

V
36.8%

time

Fig. 3.33 Decay to zero. V (t ) = V e , where V = V (0+ )

V(ss)
36.8%

V(t)

V
63.2%

time

Fig. 3.34
Stable growth from zero. V (t ) = V 1 e
V = V ( ss )

),

where

must be dimensionless (or unitless, if you will). Coefficient


A carries physical units.
Having railed on formulaic thinking, we will now present the four possible types step responses of first-order
systems formulaically, by normalizing the first-order step
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 output variable, V(t), by the range by the dynamic range, V.
The time axis is normalized by dividing time by the time
constant, . The four forms of the first-order step responses
are summarized below and shown in Figs.3.33, 3.34, 3.35,
3.36, and 3.37,
t

V (t ) = V e

where V = V (0+ )
t

(3.50)

(3.51)
V (t ) = V 1 e where V = V ( ss )
The growth and decay curves are mirror images of each
other, as shown in Fig.3.35, where the dependent variable,
V(t), has been normalized by dividing it by its maximum
value, and time is expressed in time constants. Time equal to
one time constant reduces the difference between the current value and the steady-state value by 63.2%.
The addition of a constant to the growth and decay functions, Eqs.3.50 and 3.51, offsets the growth and decay
curves vertically, Eqs.3.52 and 3.53,

3.5 First-Order Step Responses

145

Fig. 3.35a Normalized firstorder growth. b Normalized


first-order decay

v(ss)
____
=1
v(ss)

v(0)
___
=1
v(0)

v(t)
____

v(t)
____
v(0)

v(ss)

0.632

0.865

63.2%

V(t)

time

63.2%

V(0+)
0

time

Fig. 3.37 Stable exponential growth from a non-zero initial value.

) + V (0 ), where V = V (ss) V (0 )

time

time

Fig. 3.38 Stable exponential growth from a positive initial value to a


negative steady-state value

increases the energy stored in an element as an increase,


and a change that decreases the amount of energy stored as
a decrease. Consequently, we will define growth and decay
relative to the magnitude of a variable. A variable grows if
its magnitude increases. Likewise, a variable decays if its
magnitude decreases. Using this definition, it is possible to
grow from a positive initial value to a negative steady-state
value, and to decay from an initial negative value to a positive steady-state value.

 V (t ) = Ve + V( ss ) where V = V(0 ) V( ss )

36.8%

V(t)

36.8%

V(ss)

V(ss)

Fig. 3.36 Decay to a non-zero steady-state value. V (t ) = V e + V ( ss ),


where V = V (0+ ) V ( ss )

V(t) 0

V (t ) = V 1 e

63.2%

36.8%

V(ss)

time

0.865 0.950 0.982 0.993

V(0+)

V(0+)

0.950 0.982 0.993

0.632

V (t ) = V 1 e

) + V(0 )
+

(3.52)
(3.53)

Signs can create semantic problems when describing a dynamic response. For example, the plot of V(t) in Fig.3.38 is
concave upward, the shape of a decay from a positive value.
In the caption, the trace is described as a growth from an
initial positive value to a negative steady-state value. Either
description is acceptable. If the variable, V(t), was velocity and we were to observe the response of the system, we
would likely describe the response as the system reversing
direction, since the initial and final magnitudes of V(t) are
the same. It is more consistent to describe a change that

3.5.2Step Response of a First-Order Mechanical


System
We will use the force source-mass-damper system, Fig.3.13,
as the example and solve the system equations for the velocity of the mass, Eq.3.26, and force acting to accelerate the
mass, Eq.3.27, repeated below,
F0 us (t ) M dv1g
=
+ v1g
(3.26)
b
b dt
M dF0 us (t ) M dFM
=
+ FM
b
dt
b dt


for the step input, F (t ) = F0 us (t ), plotted in Fig.3.26.

(3.27)

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

146

The Heaviside step input transitions from zero to one before the response can begin. Thus, we can remove the step
function, us(t), and leave just its scaling factor, F0,
F0 M dv1g
=
+ v1g
b
b dt

M dF0 M dFM
=
+ FM
b dt
b dt

and

Our solution method is to determine the initial and final values of the response. These values are sufficient to identify
which of the four possible first-order step responses is the
correct response function.
The time constant and the final, or steady-state, value are
determined from the system equation. The time constant is
found by unit analysis after expressing the system equation
in time constant form by clearing the output variable of its
coefficient. The factor multiplying the derivative of the output variable with respect to time must have units of time for
that term to have the units of the output variable,

E sys ( 0 ) = 0 = E sys ( 0+ ) E M ( 0 ) = 0 = E M ( 0+ )
1
1
Mv12g 0 = 0 = Mv12g 0+
2
2

( )

V (t ) = V 1 e

The steady-state value is determined using the same logic


and procedure as used in the method of undetermined coefficients. It is a fact that in steady-state, all power variables
within a linear system will have the same functional form
as the input. A Heaviside step input has a single transition.
After it has transitioned from zero to one, it remains constant. Consequently, the steady state of all power variables in
a linear system driven by a step input is constant. Remember
that zero is a constant. Substitute v1g ( ss ) = C into the system
equation, and evaluate the derivative:

( )

( )

( )

M
F0
b
v1g (t ) =
1 e

v1g (t ) =

F0
b

b
t

M
1

(3.54)

The solution for force FM, acting to accelerate the mass, follows the same procedure as the solution for the response v1g.
The time constant and the steady-state value are determined
from the system equation,
M dF0 M dFM
=
+ FM
b dt
b dt

Calculation of the initial value of the output variables was


presented in Sect.3.4 and is only reviewed here. The system
was 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 the system has no energy, then
the velocity of the mass is zero.
1
v1g 0 = 0 E sys 0 = E M 0
E M 0 = Mv12g 0
2

( )

Improper ratios are difficult to read and lead to errors.


Eliminating the improper ratio in the exponent yields the
following:

F0 M dv1g ( ss )
=
+ v1g ( ss )
b
b
dt
F0
= C = v1g ( ss )
b

M
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 constant of Eq.3.26, then one or both of them would be in error.
The steady-state value of FM is determined by substituting FM ( ss ) = C into the system equation and evaluating the
derivative, as below:
M dF0 M dFM ( ss )
=
+ FM ( ss )
b dt
b
dt

( )

The energy storage variables are known as state variables.


We see that v1g is the state variable of the mass-damper system. State variables are the only variables in a system which
cannot change instantaneously,

( )

where

V = VSS

( )

v1g 0 = 0 = v1g 0+

If we know the value of the state variable at any time t, then,


also using the value of the input which we always know, we can
calculate any other power variable in the system at that time.
We can now identify the form of the step response. The
initial value of the output variable is v1g (0+ ) = 0, and the
steady-state value is v1g ( ss ) = F0 / b . The step response is a
stable exponential growth from zero of the form of Eq.3.51,

F0 M dv1g
M
=
+ v1g =
b dt
b
b

F0 M dC
+C
=
b
b dt 0

( )

M dF0
b dt

=
0

M dC
+C
b dt 0

0=C

FM ( ss ) = 0

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

F(t)

u s(t-t a )
1

No power flows into


the mass in the steadystate of a step response.
All power flows into the
damper and is dissipated
as heat.

0
g
Fig. 3.39 Linear graph of the mass-damping system showing the
power flows in steady-state

stored in the mass is changing, then the system has yet to


reach steady-state in response to the step input,
dv1g
dE
dE
= v1g M
=0
= v1g FM P = 0 FM = 0
dt
dt
dt
We can reach the same conclusion by considering the power
flows within the linear graph. If there is no change in the
amount of energy stored in the mass when the system has
reached steady-state, then the power flowing into or out of
the mass equals zero. The power flowing from the source
into the system, Psource = F (t )v1g , must all be dissipated by
shearing the lubricating fluid film, Pb = Fb v1g . In the steadystate of a step response, all power that flows into node 1 from
the source also flows out into the damping element, b. No
power flows into the mass in steady-state.
Nodes represent the value of the across variable velocity
and transmit force and power between elements. Nodes in
mechanical systems are ideally rigid and massless. Energy
is sourced, stored, or dissipated in the elements between the
nodes, not at nodes. We can express conservation of energy
in terms of power if we sum energy flow rates at nodes. The
assumed positive direction of power flows is in the direction indicated by the arrowhead on each element, Fig.3.39.
Summing power flowing into and out of node 1, where flow
in is positive and flow out is negative,

Psource = Pb + PM F0 us (t ) v1g = Fb v1g + FM v1g


When the system reaches steady-state, velocity v1g ( ss ) is
not zero. Therefore,

PM = FM v1g ( ss ) = 0 FM = 0
and
F0 us (t ) v1g = Fb v1g + FM 0 v1g

F0 us ( ss ) = F0 = Fb

Again referring to Sect.3.4, in order to determine the initial


value of FM (0+ ) at the instant after the Heaviside step transi-

time

ta

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
keeping the argument negative

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


systems state (energy storage) variable at time t = 0+ and
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
input at t = 0+ , F (0+ ) = F0 us (0+ ) = F0 . We found

( )

( )

F 0+ = b v1g 0+

( )

+ FM 0+

( )

( )

F 0+ = F0 = FM 0+

We can now identify which of the four step responses is the


correct response. The initial value is F0 = FM (0+ ), and the
steady-state value is FM ( ss ) = 0. The step response is an exponential decay to zero in the form of Eq.3.43,
t

V (t ) = V e where V = V0 FM (t ) = F0 e
b

t
(3.55)
FM (t ) = F0 e M

Physically, the mass is at rest at the instant after the Heaviside step input is applied, since its kinetic energy cannot
change instantaneously without infinite power. There is no
shear of the lubricating fluid film, since there is no motion.
Initially, all of the applied force acts to accelerate the mass.
Eventually, the velocity of the mass will create a shear force
equal to the applied force. The system will then be in steadystate. Since none of the applied force will act to accelerate
the mass, there will be no further change. All of the power
variables will have constant values.

3.6 Time Shift


Time t = 0 is arbitrary and defined by the engineer typically
at the beginning of a time period of interest. We need to address the general case when a single step input is applied to
a system but at a time other than t = 0. The step input and
the corresponding step response are formulated using a time
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
needs to be done to time shift the Heaviside unit step function, since us (t ta ) is equal to zero when (t ta ) 0.

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

148

3.7Superposition of Heaviside Step


Functions

v(ss)

v(t) 0

Combining two or more step functions allows construction


of system inputs that approximate any arbitrary function. A
pulse input is created by summing, or superposing, a step
function of positive amplitude with a step function of negative amplitude,

This portion of the


computed response
violates cause and
effect. It is before
the input acts on
the system.

ta

time

(3.58)
Unit Pulse(t , ta ) = us (t ) us (t a )

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
possible to evaluate the function for times before the input is applied,
violating cause and effect

v(ss)

v(t) 0

Mutliplying the response


function by the Heaviside
unit step with the same
time shift zeros out this
portion of the response.

ta

time

Fig. 3.42 Time shifted step response multiplied by a time shifted


Heaviside unit step function, Eq.3.57

Time shifting a response function is only slightly complicated by the fact that a time shifted response function has
a non-zero value if evaluated for time before the input that
created the response acted on the system. The result of such
a calculation is clearly nonsense, since it violates cause and
effect. However, it may not be immediately apparent from
the response plot alone that the result is nonsense, because
the plot will be a continuous exponential function. For example, the stable exponential growth, Eq.3.56, plotted in
Fig.3.41,


v1g (t ) =

b
( t ta )
F0
m
1 e

(3.56)

was shifted forward to time t=ta, but the function was plotted from time t = 0. The only way to spot the error would
be to know the correct initial value of the response function.
The step response function must be zeroed out for
(t ta ) 0, by multiplying it by a time shifted Heaviside
unit step function, Eq.3.57, Fig.3.42,


b
F
( t ta )
0

v1g (t ) = us (t ta ) 1 e m

(3.57)

The amplitude of a unit pulse is scaled by multiplying it by


a coefficient. An arbitrarily shaped function of time can be
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
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
amplitude. The first case is described as a pulse train, in
which the pulses may or may not be of equal amplitude,
duration, and time-spacing, Fig. 3.46. The second case is
usually accomplished by pulse-width-modulation, in
which the duration of the on portion of a pulse is scaled
by the desired equivalent DC or constant amplitude. Pulsewidth-modulation is useful because the energy storage elements in our systems accumulate or integrate the power
pulses from the input source, and, as a result, smooth out
variation in the systems power variables. Practical pulsewidth-modulated amplifiers operate at pulse frequencies
which are at least an order of magnitude greater than the
upper limit of human hearing.
Pulses can also be used to approximate continuous functions, such as sinusoids. Applying a concept familiar from
calculus, decreasing the duration (or width) of the pulses
increases the precision of the approximation, Fig.3.47.
Rather than reducing the width to infinitesimal, we stop at
a finite value to yield a discontinuous input function with a
stair step appearance. A trapezoidal approximation is more
accurate but limited in its applicability. Real time systems,
such as digital controllers and inverters, which approximate the sinusoids of AC current from DC power, use a
zero-order-hold which holds the variable constant until
the next pulse and creates rectangular pulses. Pulses are
the basis of digital control with the zero-order-hold using a
pulse of unit amplitude constructed as shown in Fig.3.43,
Eq.3.58.
Although we can superpose pulse responses to approximate the response of a linear model to an arbitrary input,
as shown in Fig.3.47, and then construct the response by
superposing the corresponding step pulse responses, it is not
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


Fig. 3.43a Unit step functions,
us(t) and us(tta). b Sum or
superposition of the two unit step
functions create a unit pulse of
width ta

149

us(t)

us(t)

us(t) - us(t-ta)
0

ta

time

-1

time

ta

-1

-us(t-ta)
Fig. 3.44a Force pulse,
F(t) = 10N us(t) + 10N us(t1)
20N us(t2).
b Force steps superposed to create
the pulse

-us(t-ta)

20

20
10 us(t)
10

10

F(t) 0
0
N

t, sec

-10

-5

t, sec

-20 us(t-2)

-20

10

F(t) 0
N

-10

-20

Fig. 3.45a Force pulse,


F(t) = 10N us(t)15N us(t1)+5N us(t3).
b Force steps superposed to create the
pulse

10 us(t-1)

F(t) 0
0
N

10 N u s(t)
10
5 N u s(t-3)

5
1

t, sec

F(t) 0
N

-5

-10

-10

-15

-15

system equations to arbitrary inputs, after we introduce the


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

Linear equations can be viewed as linear operators since


they operate on the input variable to yield the output variable. Our differential system equations are linear operators.
Linear operators have two valuable properties, scaling and

t, sec

15 N us(t-1)

Pulse(t)

3.8Superposition of First-Order Step


Responses

Period T

0.5T

1.5T

time

2T

2.5T

Fig. 3.46 Unit pulse train with 50% duty cycle. Duty cycle is the percentage of a period, T, during which the pulse has non-zero amplitude

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

150

Fig. 3.47a Arbitrary input


approximated by one second
duration pulses. b Input more
accurately approximated by half
second duration pulses

F(t)
N

b
40

F(t)
N

20
0

Fig. 3.48a Example massdamper system driven by a force


source. b Pulse input of force

t, sec

20
0

Mass
M

Lubricating
fluid film
Damping b

F0

ta

t, sec

time shifted step input is Eq.3.57. Both equations are repeated below,

F0 us(t)

F0

F(t), N

F(t), N

t, sec

x,v

F(t)

40

b
t
F
(3.54)
v1g (t ) = 0 1 e M
b

ta

t, sec


-F0

-F0 us(t-ta)
Fig. 3.49 Pulse input of force plotted in Fig.3.48b created by superposing two step inputs, Eq.3.59

superposition. We can calculate the response of a system to


an input of unit amplitude, and then scale the unit input response, to calculate the response of the system to an input of
any magnitude. We can use superposition to calculate the response of a system to combined inputs or inputs constructed
from simpler functions, specifically step functions, by summing the responses to each individual input.
Example of Superposition, Response to a Pulse Input.
Calculate the response of the mass-damper system of the
previous example to a pulse input of force of magnitude F0,
as shown in Fig.3.48.
We first construct the pulse input from the superposition
of two step inputs, Fig.3.49, Eq.3.59,
F (t ) = F0 us (t ) F0 us (t ta ) = F0 us (t ) us (t ta )
(3.59)
We then superpose (sum) the response to each of the inputs,
which we have already calculated. The system to a step input
of force, F0, is Eq.3.54 and the response of the system to a

b
F
( t ta )
0

v1g (t ) = us (t ta ) 1 e m

(3.57)

It must be emphasized that the output or response function


has the same time shift as the input function which created it,
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 subtracting time ta from the time variable t to synchronize the
response function with the input function. The time shifted
response function must be multiplied by a Heaviside step
function with the same time shift, in order to zero-out the
result of calculating the response, prior to the time the corresponding input acts on the system.
The analytical expression of the superposition of step responses is a straightforward sum, Figs. 3.51 and 3.52. It is
important to remember that time shifted step response functions must be multiplied by time shifted unit step functions,
in order to zero-out (or null) the functions for time, t < ta,


b
t
F
v(t ) = us (t ) 0 1 e m
b

b
(t ta )
F
+ us (t ta ) 0 1 e m
(3.60)
b

A Heaviside unit step function is used to zero-out the response function term, until its corresponding input is applied. Again, if the Heaviside unit step function is not used as

3.8 Superposition of First-Order Step Responses

151

Response to positive
force step input

F0
b

F0
F(t), N

vb(t), m
sec

ta

time

-F0

-F0
Response to negative b
force step input

stant of a linear system is independent of the amplitudes of


the input and output variables. If this seems counter-intuitive,
it is because the timescales of the responses of most systems
we are familiar with in everyday life depend on the amplitude
of the output. A system with a time constant (which is not
constant) is non-linear. For example, drop a coin on a hard
surface, and listen to it, as it rattles to rest. The frequency increases, as the rattle amplitude decreases. If the rattling coin
were described by a linear system equation, it would have a
constant frequency. In contrast, a non-linear system does not
have a single oscillation frequency. It has a range of frequencies. The timescaling of a non-linear systems 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)

Superposition (sum) of
the two step responses

Response to positive
force step input

F0
b

v1g(t)
m
sec
ta

t, sec

Response to negative
force step input

-F0
b

Fig. 3.51 Superposition of the positive and negative step responses


showing the summation

v1g(t)
m
___
sec
0

ta

t, sec

Fig. 3.52 Pulse response. Growth during the pulse from t = 0 to t = ta,
followed by decay after time, t = ta

3.8.1Scaling, Time shifting, and Superposing


Unit Step Responses
A particularly efficient method of using superposition to calculate the response of a first-order linear system to an arbitrary pulse input is as follows:
1. Construct the input function by scaling, time shifting, and
superposing (summing) Heaviside step functions.
2. Assume the system is de-energized, and determine the
response of the output variable to an input step of unit
magnitude.
3. Scale, time shift, and superpose (sum) the unit step response, using the same time shifting and scaling, as used
to construct the input pulse from Heaviside step functions. Multiply each term in the response function by the
same time shifted Heaviside step function used, to create
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 responses, in order to construct the response function, are possible only if we restrict models of energetic attributes of the
system to linear models, as we have done.
We will illustrate the technique using the force sourcespring-damper system of Sect.3.2.4.1, Fig.3.54. The energetic equations and the system equations for the force
through the spring, FK, Eq.3.29, and the force through the
damper, Fb, Eq.3.30, are reproduced below for convenience.
Energetic Equations

a multiplicative factor with the corresponding step response


function term of the same time shift, then the result includes
the evaluation of the step response for the entire period of
calculation, rendering said result non-causal for its violation of cause and effect.
Both step responses have the same time scaling, because
the time constant of a system depends on the systems parameters, and how the elements are interconnected. The time con-

Continuity or Node Eq: F (t ) Fb FK = 0


Compatibility or Path Eq: v1g = v1g
Element Eqs: Fb = bv1g
Energy Eqs: E sys = E K

dFK
= Kv1g
dt

EK =

FK2
2K

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

152
Fig. 3.53 Block diagram
showing the pulse inputpulse
response, as an inputoutput
relationship

F(t)

v1g(t)

time

Pulse Input

Fig. 3.54a Energetic schematic


of a force source acting on an
ideal rigid and massless bar,
which is connected to a spring,
K, and dashpot, b. b Linear graph
of the system shown schematically in a

Linear
System

time

Pulse Response

x,v

K
F(t)

F(t)

20

Rigid and massless bar


Fig. 3.55a Input force pulse.
b Scaled and time shifted
Heaviside step functions which
superpose to create the pulse in a

20
10 us(t)
10

10
F(t)
0
N
0

F(t)
N

3
t, sec

-10

-10

-20

-20

10N us(t-2)
0

3 t, sec

-20N us(t-1)

F (t ) =

b dFK
+ FK
K dt

(3.29)

b dF (t ) b dFb
(3.30)
=
+ Fb
K dt
K dt
We will solve these system equations for the force input in
the force pulse, shown in Fig.3.55a, which is created by
scaling and superposing Heaviside unit step functions, as
shown in Fig.3.55b.
The input function is


F (t ) = 10 N us (t ) 20 N us (t 1) + 10 N us (t 2)

(3.61)

To use superposition, the unit step response must be derived


by assuming that the system is de-energized, even if the system is clearly not de-energized. The energy in the system

from earlier inputs will be accounted for by response functions time shifted to those earlier times.
We begin by finding the unit step response of the force
acting through the spring, FK. We can establish the time constant and the steady-state value of the output variable from
the system equation in time constant form,
F (t ) =

b dFK
b
+ FK =
K dt
K

The system is in steady-state when all of the power variables


in the system have the functional form of the input, which for
a step input is constant in steady-state. Hence, the derivative
of FK is zero in steady-state, since the spring force is constant:
1 N u s ( ss ) =
1

b dFK ( ss )
+ FK ( ss )
K
dt
0

1 N = FK ( ss )

3.8 Superposition of First-Order Step Responses

153

10

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

E sys ( 0 ) = 0 = E K ( 0 ) E K ( 0 ) =

( )F

FK2 0

2K

-10

( )

We identify the form of the first-order step response from


FK (0+ ) = 0 and FK ( ss ) = 1 N as a stable exponential growth
from zero with the form of Eq.3.53 and Fig.3.34,

FK (t ) = 1N 1 e
u .s.

FK (t ) = 1N 1 e
u .s.

K
t
b

)10 u (t ) 1N (1 e ) 20 u (t 1)
)10 u (t 2)
+ 1N (1 e

FK (t ) = 1N 1 e

K
(t 1)
b

K
(t 2 )
b

b dF (t ) b dFb
b
=
+ Fb time constant =
K dt
K
dt
K

t, sec

There is only one time constant in a first-order system. If


we had arrived at a different time constant than that for the
spring force in the same system, one of the two (or both)
of the time constants would be erroneous. In steady-state,
the output variable must have the form of the input variable,
which, in steady-state, is a constant,
b d1N us ( ss )
K
dt

=
0

b dFb ( ss )
+ Fb ( ss )
K
dt
0

Fb ( ss ) = 0

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


input and the state (energy storage) variable, FK, at time
t = 0+ and the energetic equations. We know F (0+ ) = 1N and
FK (0+ ) = 0. We need Fb (0+ ). Using continuity,

( )

( )

( )

( )

( )

( )

F 0+ Fb 0+ FK 0+ = 0 F 0+ FK 0+ = Fb 0+

( )

Fb 0+ = 1 N
Comparing Fb (0+ ) = 1 N with Fb ( ss ) = 0, we identify the
unit step response function as a decay to zero in the form, as
expressed in Eq.3.50 and Fig.3.33,
t

Fb (t ) = 1N e

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


b = 2,500 N sec/m to evaluate the response, plotted in
Fig.3.56.
We will now repeat the process to determine the response
of the damper force, Fb, to the force pulse input. We find the
time constant and the steady-state response to a step input of
1N from the system equation,

Fig. 3.56 Response function, FK(t), evaluated using K = 5,000 N/m


and b = 2,500 N sec/m

We construct the pulse response function by weighting (scaling) the unit step response function with the same scaling
factors and time shifts used to construct the input function.
Both the exponents of the exponentials and the Heaviside
unit step function must be time shifted. Each term in the
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
response function,
K
t
b

-5

FK 0 = 0 = FK 0+

FK(t), N 0

(0 ) = 0

A state (energy storage) variable cannot change instantaneously without an infinite power flow to transfer a finite
amount of power during an infinitesimal time. Consequently,
the state variable, FK, has the same value at time, t = 0 , before the input step is applied as it does at time, t = 0+, at the
instant after the step is applied,

( )

u .s.

Fb (t ) = 1N e

K
t
b

u .s.

We construct the pulse response function by weighting (scaling) the unit step response function with the same scaling
factors and time shifts used to construct the input function,
Eq.3.61, F (t ) = 10 N us (t ) 20 N us (t 1) + 10 N us (t 2),


Fb (t ) = 1Ne

K
t
b

10us (t ) 1Ne

+ 1Ne

K
(t 2 )
b

K
(t 1)
b

10us (t 2)

20us (t 1)

(3.62)

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

154

10

Fb (t), N
-10

-20
0

t, sec

Fig. 3.57 Response function, Fb(t), evaluated using K = 5,000 N/m


and b = 2,500 N sec/m

The damper force picks up the instantaneous changes of the


input force, which are then transferred to the spring as the
spring deforms (Fig.3.57).

3.9Superposition of Second-Order Step


Responses
We will demonstrate the superposition of second-order step
responses using the force source-spring-mass-damper system
of Sect.3.2.4.3, shown as an energetic schematic and a linear
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.
Energetic Equations

Compatibility or Path Eq: v1g = v1g


dFK
= Kv1g
dt

Energy Eqs: E sys = E K + E M




EK =

FM = M
FK2
2K

EM =

3.9.1Overdamped or Non-Oscillatory Step


Response
We will use the parameter values K = 5, 000 N/m,
b = 2,500 N sec/m, and M = 200 kg for the overdamped,
non-oscillatory case. Substituting these values into the system equation for the force acting through the spring, Eq.3.33,

Continuity or Node Eq: F (t ) = Fb + FK + FM

Element Eqs: Fb = bv1g

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
values. Second-order step responses are a continuum of responses, from those which resemble the step response of a
first-order system, to those which are almost sustained oscillations. The increased variation in potential responses is
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
depends on the relative energy storage capacity of the two
storage elements, and on the rate at which energy is dissipated as heat, during the flow of energy between the two
storage elements. If the eigenvalues (characteristic values)
of the system are purely real, then the step response is overdamped or non-oscillatory. Conversely, if the eigenvalues
are complex conjugates, then the step response is underdamped or oscillatory. Oscillatory is in quotation marks,
because one normally thinks of an oscillation as consisting
of at least one complete cycle. An underdamped systems
step response will overshoot or undershoot the final steadystate value, but it may not complete an entire cycle.
We will solve the system equation for the force acting
through the spring, Eq.3.33, using the method of undetermined coefficients first for parameter values that yield an
overdamped or non-oscillatory response. We shall then do
likewise for parameter values which yield an underdamped
or oscillatory response.

dv1g

d 2 FK
K
b dFK K
F (t ) =
FK
+
+
2
M
M dt
M
dt

dt
1
Mv12g
2

N
N sec
N
2,500
5, 000
2
d
F
dF
m F (t ) =
K
m
K
mF
+
+
K
200 kg
200 kg
200 kg
dt
dt 2

5, 000

d 2 FK
K
b dFK K
F (t ) =
FK
+
+
2
M
M dt
M
dt

(3.33)

d 2 FK 12.5 dFK
25
25
=
+
+
F
t
FK
(
)
2
2
sec dt
sec
dt
sec 2

2
1 dF (t ) d v1g b dv1g K
=
+
+ v1g
M dt
M dt
M
dt 2

(3.34)

Particular Solution We will determine the particular solution for a unit step input of force, F (t ) = 1N us (t ). The
particular solution is the steady-state solution, after the homogeneous response has decayed to zero. The steady-state
response for all power variables in a system is the functional

The methodology differs from the superposition of firstorder step responses only in the manner that the unit step

3.9 Superposition of Second-Order Step Responses


Fig. 3.31a Force source springmass-damper system of Fig.3.23.
b Fig. 3.24b

155

b
x,v

F(t)

F(t)

Lubricating fluid
Damping b
form of the input variable. If the input is a Heaviside step,
then in steady-state all power variables will be constants,

General Solution Sum the particular and homogeneous solutions to form the general solution:

d 2 FK ( ss )
25
12.5 dFK ( ss )
25
1N
u
ss
=
+
+
FK ( ss )
)
s(
2
sec
dt
sec 2
sec
dt 2
0
0

FK (t ) = FK P ( ss ) + FK H (t ) FK (t ) = 1N+A1e s1t + A2 e s2t

FK P ( ss ) = 1 N.
Homogeneous Solution Form the homogeneous equation
by setting the input to zero:
0=

d 2 FK 12.5 dFK
25
FK
+
+
2
sec dt
sec 2
dt

Perform the Laplace transformation (neglecting initial condition terms) to form the characteristic equation. We will
now drop the units of sec.

L {0} = L

d 2 FK

dFK
+ 25 FK
2 + 12.5
dt
dt

0 = s 2 FK ( s ) + 12.5sFK ( s ) + 25 FK ( s )

Use the initial conditions to determine the coefficients A1


and A2. In order to use superposition, we must assume that
the systemis de-energized prior to the a pplication of the first
input. The initial conditions needed to solve this equation are
the values of the output variable, FK (0+ ), and its derivative,
dFK (0+ ) /dt , at time t = 0+ , the instant after the step input
is applied to the system. The output variable, FK, is a state
(energy storage) variable. If the system is de-energized at
time t = 0 , the energy storage variables equal zero. The state
variables cannot change instantaneously, since time is needed to move energy into or out of the energy storage elements.
Therefore, the state variables remain zero at time t = 0+ , and

( )

FK 0+ = 0
In Sect.3.4.2, we established

( ) = Kv

dFK 0+

0 = s + 12.5s + 25 FK ( s )
2

0 = s 2 + 12.5s + 25
Solve the characteristic equation to determine eigenvalues
of the system:
s1 , s2 =

12.5 12.52 4.25


2

s1 , s2 = 6.25 3.75 s1 = 2.5, s2 = 10


The eigenvalues are real. The system is overdamped, and the
step response is non-oscillatory.
Form the homogeneous solution
FK H (t ) = A1e s1t + A2 e s2t

1g

dt

(0 )
+

The velocity of the mass, v1g, is the other state variable of the
system. Hence, v1g (0+ ) = 0 and

( ) = Kv

dFK 0+
dt

1g

(0 ) = 0
+

Apply the initial condition, FK (0+ ) = 0 , to the general


solution:

( )

FK 0+ = 0

0 = 1N + A1 e s1 0 1 + A 2 e s2 0

0 = 1N + A1 + A 2

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

156

dFK (t )
= 0, to the general
dt

Apply the initial condition,


solution:
dFK (t )
dt

d
1 N+A1e s1t + A 2 e s2t
dt

1.0

)
FK(t), N 0.5

dFK (t ) d1N
de s1t
de s2t
+ A1
+ A2
=
dt
dt 0
dt
dt
dFK (t )

= A1e s t

ds1t
ds t
+ A2 es t 2
dt
dt

= A1e s t s1

dt
dt
+ A 2 e s t s2
dt
dt 0

dt
dFK (t )
dt

dFK (t )
dt

= A1e s1t s1 + A 2 e s2t s2

Evaluate for t = 0+ :
dFK (t )
=0
dt

0 = A1 e s1 0 1 s1 + A 2 e s2 0 1 s2

0 = A1 s1 + A 2 s2
Form a system of two equations in the unknowns A1 and A2.
Solve the system for A1 and A2. We will use direct elimination:
0 = s1 A1 + s2 A2
s
s
A1 = 2 A2 1N = 2 A2 + A2
1N = A1 + A2
s1
s1
s

1N = 2 + 1 A2
s1
1N =

A1 =

s1 s2
A2
s1

s1
1N
s1 s2

s2
s s1
s2
A2 A1 = 2
1N A1 =
1N
s1
s1 s1 s2
s1 s2

s2
10
1N A1 =
1N A1 = 1.33 N
2.5 ( 10)
s1 s2

A2 =

( 2.5)
s1
1N A2 =
1N A2 = 0.333N
s1 s2
2.5 ( 10)

Substitute into the general solution:


FK (t ) = 1N+A1e s1t + A2 e s2t
FK (t ) = 1N 1.33Ne 2.5t + 0.333Ne 10t

t, sec

This is the unit step response and is plotted in Fig.3.58.


The overdamped second-order unit step response,
Fig.3.58, strongly resembles the first-order stable exponential growth, shown in Fig.3.34. A second-order step response
differs from a first-order step response by the reverse
curve at the base or beginning, seen in both the unit step
response and the pulse response of the overdamped secondorder spring-mass-damper system.
We now construct the pulse response function by weighting (scaling) the unit step response function,
FK (t ) = 1N 1.33Ne 2.5t + 0.333Ne 10t
with the same scaling factors and time shifts used to construct the input function, Eq.3.61,
F (t ) = 10N us (t ) 20N us (t 1) + 10N us (t 2)
The most common error is to omit the time shift in one or
more of the exponential terms when constructing the response function, Fig.3.59:

FK (t ) = 1N 1.33Ne 2.5t + 0.333Ne 10t 10us (t )

(
+ (1N 1.33Ne
1N 1.33Ne

Substitute for the eigenvalues, s1 = 2.5, s2 = 10:


A1 =

Fig. 3.58 Unit step response function, FK(t), evaluated using


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

s
s
1N = 2 + 1 A2
s1 s1
A2 =

)
)10u (t 2)

2.5(t 1)

+ 0.333Ne 10(t 1) 20us (t 1)

2.5(t 2)

10 (t 2)

+ 0.333Ne

3.9.2Underdamped or Oscillatory Step


Response
N
use the parameter values, K = 5, 000 ,
m
N sec
b = 2,500
, and M = 400 kg, for the underdamped,
m
oscillatory case. Substitute these values into the system
equation for the force acting through the spring, Eq.3.33,

We

will

d 2 FK
K
b dFK K
F (t ) =
FK
+
+
M
M dt
M
dt 2

3.9 Superposition of Second-Order Step Responses

157

10

Increasing the mass from 200kg to 400kg changed the eigenvalues from a real pair to complex conjugates. When we
express the quadratic equation in terms of the systems parameters,

FK(t), N 0

d 2 FK
K
b dFK K
+
+
F (t ) =
FK
2
M
M dt
M
dt

-5

-10

t, sec

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

N
N sec
N
2,500
5, 000
2
d
F
dF
m F (t ) =
K
m
K
mF
+
+
K
400 kg
400 kg
400 kg
dt
dt 2

5, 000

d 2 FK 6.25 dFK 12.5


12.5
F (t ) =
FK
+
+
2
sec dt
sec
sec 2
dt 2
Particular Solution Determine the particular solution for a
unit step input of force, F (t ) = 1N us (t ) :
d 2 FK P ( ss )
12.5
6.25 dFK P ( ss )
12.5
1N
u
ss
=
+
+
FK ( ss )
(
)
s
2
2
sec
dt
dt
sec
sec 2 P
0
0
FK P ( ss ) = 1N

Homogeneous Solution Form the homogeneous equation


by setting the input to zero:
0=

d 2 FK 6.25 dFK 12.5


+
+
FK
sec dt
dt 2
sec 2

Perform the Laplace transformation (neglecting the initial


condition terms) and form the characteristic equation. We
will now drop the units of sec.

L {0} = L

d 2 FK

dFK
+ 12.5 FK
2 + 6.25
dt
dt

0 = s FK ( s ) + 6.25sFK ( s ) + 12.5 FK ( s )
2

b
K
b
4
M
M
M
s1 , s2 =
2
we see that increasing mass M reduced both the ratio of the
damping coefficient b to mass M and ratio of the spring constant K to mass M.
The ratio between damping and mass is due to the physical
mechanisms of energy dissipation and kinetic energy storage.
Energy is dissipated as heat due to shear motion. Increasing
the mass of the system reduces the velocity v1g needed to store
an amount of energy, decreasing the rate of kinetic energy lost
as heat. For an oscillation to occur, the energy stored in one
independent energy storage mode must flow into to the other
mode and back again. The ratio of the spring constant K over
mass M governs the relative magnitudes of the energy storage (state) variables of the system and the systems ideal, undamped, natural frequency, n. Increasing the spring constant
K allows the systems energy to be transferred from kinetic
to elastic strain energy with less displacement of the mass,
decreasing the time needed to transfer kinetic to strain energy,
and, thereby, increasing the frequency of the oscillation.
We now form the homogeneous solution,
FK H (t ) = A1e s1t + A2 e s2t
The symbolic form of the homogeneous solution is the same
for all second-order systems. The solutions differ in the values of the eigenvalues and the undetermined coefficients.
General Solution Sum the particular and homogeneous
solutions for the general solution:
FK (t ) = FK P ( ss ) + FK H (t ) FK (t ) = 1N+A1e s1t + A2 e s2t
The general solution in symbolic form is also identical to
that of the overdamped case. We established the initial conditions in Sect.3.9.1 to be the following:

0 = s 2 + 6.25s + 12.5 FK ( s )

FK (t ) = 0 and

0 = s + 6.25s + 12.5
Solve the characteristic equation to determine the eigenvalues of the system:
s1 , s2 =

6.25 6.252 4 12.5


s1 , s2 = 3.13 j1.65
2

dFK (t )
=0
dt

We applied the initial conditions to the general solution of


the overdamped case and found the following:
A1 =

s2
1N and
s1 s2

A2 =

s1
1N
s1 s2

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

158

These results for A1 and A2 apply for all eigenvalues s1


and s2. We shall see that the overdamped and the underdamped solutions are quite different because their eigenvalues differ. The eigenvalues of this underdamped case are
s1 , s2 = 3.13 j1.65.
We will defer substituting in numerical values as long
as possible. Our first substitutions will be s1 = + jw and
s2 = jw ,
A1 =

s2
1N
s1 s2

A1 =

A1 =
A2 =

s1
1N
s1 s2

A2 =

( + jw )
1N
j 2w

FK (t ) = 1N + A1e + A2 e

s2 t

( + jw )
jw
1N e t + jwt +
1N e t jwt
j 2w
j 2w

Use the property of exponentials, e a + b = e a eb :

jw
( + jw )
1Ne t e jwt +
1Ne t e jwt
j 2w
j 2w

t
Factor out 1Ne :

FK (t ) = 1N +

-2

.54

1= 2.66 rad

-1

2
-j
-j1.65
-j2

Fig. 3.60 Eigenvalues, s1 = 3.13 + j1.65 and s2 = 3.13 j1.65, plotted as vectors in the s-plane, s = + jw

Express the eigenvalues as complex exponentials by calculating their angle (arguments) and magnitude (modulus). Always sketch complex numbers to ensure that their angles are
calculated correctly, Fig.3.60,
s1 = 3.13 + j1.65 = 3.54 e j 2.66
and
s1 = 3.13 j1.65 = 3.54 e j 2.66
FK (t ) = 1N +

1Ne 3.13t 3.54e j 2.66 j1.65t 3.54e j 2.66 j1.65t


e
e

j2
j2
1.65

Factor out a magnitude of 3.54. Place the remaining terms


over the common denominator, j2:
FK (t ) = 1N +

3.54 N e 3.13t e j 2.66 e j1.65t e j 2.66 e j1.65t

1.65
j2

Use the property of exponentials, e a eb = e a + b:


e j1.65t j 2.66 e j1.65t + j 2.66
FK (t ) = 1N + 2.15 N e 3.13t

j2

1Ne t jw jwt + jw jwt


e
e
w j 2
j2

The parenthetical quantity resembles the form we need to


use Eulers sine formula, Eq.2.54. We now substitute in values for the eigenvalues, s1 = + jw , and s2 = jw ,
s1 = 3.13 + j1.65

and

s2 = 3.13 j1.65

FK (t ) = 1N
+

-3.13

( + jw )
jw
1N e( + jw )t +
1N e( jw )t
j 2w
j 2w

Distribute time t into the exponents:

FK (t ) = 1N +

-4

s-plane
j1.65

=3

s 2 = -3.13 - j1.65

( + jw )
1N
+ jw ( jw )

s1t

FK (t ) = 1N +

jw
1N
+ jw ( jw )

Substitute these expressions for A1 and A2 and the symbolic


form of the eigenvalues into the general solution:

FK (t ) = 1N +

|s |

jw
1N
j 2w
A2 =

s1 = -3.13 + j1.65

1N e 3.13t 3.13 j1.65 j1.65t 3.13 + j1.65 j1.65t

e
e

1.65
j2
j2

Factor the imaginary number j out of the exponents:


e j (1.65t 2.66) e j (1.65t 2.66)
FK (t ) = 1N + 2.15 N e 3.13t

j2

Use Eulers sine formula, Eq.2.54, to eliminate the complex


numbers,
e j e j
(2.54)
= sin ( )
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


a unit step input, where force, F (t ) = 1N us (t ), remains unchanged, since it is the steady-state response:

1.0

12.5
1N us ( ss ) =
sec 2

FK(t), N 0.5

d 2 FK P ( ss )
dt

t, sec

This is the unit step response, plotted in Fig.3.61. This underdamped response appears overdamped. The response
overshoots the steady-state value by a mere 0.1%. It does
not oscillate; yet the unit step response has a sine factor with
a frequency of 1.65rad/sec.
The lack of oscillation is due to the relative magnitudes of
the real and imaginary components of the eigenvalues. The
decay rate is the real component sigma, = 3.13. Sigma is
approximately twice the magnitude of the angular frequency,
w = 1.65 rad/sec. In the time necessary to complete an oscillation, the period as T,
2

2
T=
=
= 3.81sec
w 1.65 rad
sec

d 2 FK
K
b dFK K
F (t ) =
FK
+
+
2
M
M dt
M
dt
N
N sec
N
500
5, 000
2
d
F
dF
m F (t ) =
K
m
K
mF
+
+
K
400 kg
400 kg
400 kg
dt
dt 2

5, 000

6.25 dFK P ( ss )
12.5
+
FK ( ss )
sec
dt
sec 2 P
0

Homogeneous Solution Form the homogeneous equation.


Drop the units of sec. Perform the Laplace transformation,
neglecting initial condition terms. Set the characteristic function to zero to yield the characteristic equation:
0 = s 2 + 1.25s + 12.5
Solve the characteristic equation to yield the eigenvalues:
s1 , s2 =

1.25 1.252 4 12.5


s1 , s2 = 0.625 j 3.48
2

The time constant of the decay envelop described by these


eigenvalues is
1

1
= 1.6 sec
0.625

= deacy

and the period of the oscillation is


2

the real exponential decay, e , has progressed to


e 3.13T = e( 3.13)(3.81) = 6.62 10 6 . The energy in the system is
dissipated too quickly for the system to complete a cycle of
oscillation.
We will reduce the damping in the system from
b = 2,500 N sec/m to b = 500 N sec/m , while keeping
K = 5, 000 N/m and M = 400 kg, in order to create a more
oscillatory underdamped response. Substitute these values
into the system equation for the force acting through the
spring, Eq.3.33, as shown below:

FK P ( ss ) = 1N

Fig. 3.61 Underdamped unit step response. The eigenvalues are


s1 , s2 = 3.13 j1.65. The response overshoots approximately 0.1% of
the steady-state value

=T

2
= 1.8 sec
rad
3.48
sec

which will allow the oscillation to persist for five cycles.


Again, the homogeneous solution of all second-order systems is the following:
FK H (t ) = A1e s1t + A2 e s2t
General Solution The general solution is also identical to
the previous one:
FK (t ) = 1N + A1e s1t + A 2 e s2t
Apply the Initial Conditions. We find the same form of the
undetermined coefficients, A1 and A2, expressed in terms of
the eigenvalues written symbolically as s1 , s2 = jw ,
A1 =

jw
1N
j 2w

and

A2 =

( + jw )
1N
j 2w

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

160

j
s 1 = -0.625+ j3.48

j4

s-plane

1.5

j3.48
j3

|s1| = 3.54

j2

-2

1.0

FK(t), N
0.5
0

1= 1.77 rad
1

-1 -0.625

Fig. 3.62 The eigenvalue, s1 = 0.625 + j 3.48, plotted as a vector in


the s-plane, s = + jw

Substitute these expressions for A1 and A2, as well as the


symbolic form of the eigenvalues into the general solution:
FK (t ) = 1N +

( + jw )
jw
1N e( + jw )t +
1N e( jw )t
j 2w
j 2w

Distribute time, t, into the exponents, using the property of


1Ne t
exponentials, e a + b = e a eb . Factor out
, and place over
w
a common denominator to yield the following:
FK (t ) = 1N +

1N e

( jw ) e

jw t

( + jw ) e

jw t

and

s2 = 0.625 j 3.48

FK (t ) = 1N +

s1 = s2 = 0.625 + j 3.48 =

t, sec

10

Fig. 3.63 Underdamped unit step response. The eigenvalues are


s1 , s2 = 0.625 j 3.48

Substitute into the unit step response:


FK (t ) = 1N +

1N e 0.625t 3.54 e j1.75 e j 3.48t 3.54 e j1.75 e j 3.48t


3.48
j2

Factor out the magnitude 3.54 and use the property of exponentials, e a + eb = e a + b:
FK (t ) = 1 N +

3.54 N e 0.625t e j 3.48t j1.75 e j 3.48t + j1.75


3.48
j2

FK (t ) = 1N + 1.02 N e 0.625t

e j (3.48t 1.75) e ( j 3.48t j1.75)


j2

We now have the proper form to use Eulers sine formula,


sin( ) =

e j e j
j2

e j (3.48t 1.75) e ( j 3.48t j1.75)


= sin(3.48t 1.75)
j2

j 3.48t
( 0.625 + j 3.48) e j 3.48t
1N e 0.625t ( 0.625 j 3.48) e
3.48
j2

Express the eigenvalues as complex exponentials. Calculate


the magnitude (modulus) and angle (argument), Fig.3.62:

( 0.625)2 + 3.482

= 3.54

3.48
The angle (argument) is tan 1
= 1.75 rad,
0.625
s1 = 0.625 + j 3.48 = 3.54e j1.75
and

j2

We now substitute in values for the eigenvalues s1 = + jw


and s2 = jw :
s1 = 0.625 + j 3.48

s2 = 0.625 j 3.48 = 3.54e j1.75

yielding the purely real unit step response function, FK (t ),


u .s.
below, and plotted in Fig.3.63,
FK (t ) = 1N + 1.02 N e 0.625t sin (3.48t 1.75)
u .s.

Construct the pulse response function by superposing unit


step responses which are scaled and time shifted with the
same scaling and time shifting as the input function. The
input function, F(t), Eq.3.61, shown in Fig.3.55, is
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

FK(t), N

F2 = 250

0
F1 = 100

-10

-20
-30

t, sec

Fig. 3.65 Representation of a previously applied step input. The zigzags indicate discontinuities in the time axis

10

Fig. 3.64
Response of the spring force, FK(t), in the underdamped spring-mass-damper system with the eigenvalues,
s1, s2 = 0.625 j 3.48 , to the pulse input shown in Fig.3.55

the use of zigzags on the horizontal lines, representing F1 (t )


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.
We can calculate the response of the system to the step
input applied at time t = 0 using superposition. The input

All arguments in each term of the pulse response function


must be time shifted by the same amount,

time

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.

3.10 Initial Condition of Energized Systems


Our previous calculations assumed that the system was deenergized, prior to the application of a step input. This was
also true, when we used the superposition of multiple step
inputs to create pulses, or to approximate an arbitrary input.
We assumed that the system was de-energized, prior to the
application of the first step input, and we assumed that all
corresponding superposed step responses started from a deenergized state. Otherwise, superposition cannot be used.
The remaining case to consider is that of a system energized
before a step input is applied to it at time t=0. It is common for a system to be energized and in already steady-state
under a step input applied previously, at an unknown time,
when a new step input is applied to the system. For example,
a DC electric motor may be running in steady-state under an
applied voltage of 12 VDC, when a step change increases the
voltage to 36 VDC.
We will use the force source-mass-damper system shown
in Figs.3.13 and 3.48 as the example system. The system is
in steady-state under a step input of 100N when we increase
the applied force to 250N, as shown in Fig.3.65. We define
time, t = 0, for our convenience, as the time when the second
step input of force, F us (t ) = 150 N us (t ), is applied. Note

function, F(t), can be constructed by superposing (summing) two step inputs but we must first estimate the minimum duration before time t = 0 the first input could have
been applied such that the system reached steady-state by
time t = 0.
The largest time constant in a system, max, or, equivalently, the smallest magnitude real eigenvalue component,
min, in the system determines the duration of the transient
period of a step response. Although mathematically, an
exponential decay never reaches zero, in reality, physical
systems reach steady-state once their current value is indistinguishable from the final value. When a system reaches
steady-state depends on the precision of the measurement,
as well as the noise and ripple in the response signal, which
are unknown before a measurement is made. When the duration of the transient period is estimated, it is calculated in
units of the largest time constant, max. Typically, the transient period of a response is estimated as five time constants, which leaves only 0.7% of the initial range of the
response remaining, Fig.3.66.
Since the mass-damper system is described as in steadystate at time t = 0 under the input force of 100N, we will
estimate minimum duration t prior to t = 0 that the first
step input was applied as t = 5 max , ta 5 max. Having set
a limit, or a constraint, on time ta we can now construct the
input function by superposing (summing) two steps. Recall

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

162

F(t), N

v(0)
____
v(0) =1

Step Input 1

13.5%

Step Input 2

F1(t) = 100 N us(t-t a )

36.8%

v(t)
_____
v(0)

F2(t) = 150 N us(t)

5.0%

1.8%

0.7%

max 2max 3max 4max 5max

t = 0-

t = 0+
0

ta

t, sec

Fig. 3.67 Input function F(t), Eq.3.64

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 argument equals zero. Be careful with signs since ta is negative. The input force F(t) is the superposition of a 100N step
which transitions at time t = ta and a 150N step which transitions at time t = 0, Eq.3.63,

v1g(t)
0.4
m
___
sec 0.2

(3.63)
F (t ) = 100 N us (t ta ) + 150 N us (t )

-0.2

0
-1

-0.5

t, sec

0.5

We will use the parameter values b = 300 N sec/m and


M = 40 kg and solve the system equation for the velocity of
the mass, Eq.3.26, repeated below,

Fig. 3.68 Response function v1g(t), Eq.3.65

F0 us (t ) M dv1g
(3.26)
=
+ v1g
b
b dt

with the same time shifting and scaling using in the input
function, Eq.3.64,

There is only one time constant in a first-order system.


Hence,

max = =

M
max =
b

v1g (t ) =

40 kg
max = 0.13sec
N sec
300
m

We shall set the time of the previous step input to seven times
the time constant prior to t=0,
ta = 7 max ta = ( 7 )( 0.13) sec = 0.91sec

( t + 0.91)

100 N
0.13
1
e

us (t + 0.91sec )
N sec
300
m

150 N
0.13
1
e

us (t )
N sec
300
m

v1g (t ) = 0.33

( t + 0.91)

m
0.13
1
e

us (t + 0.91sec )
sec

+ 0.5
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.
We have already found the step response v1g(t), Eq.3.54,
repeated below,
b
t
F
v1g (t ) = 0 1 e M
(3.54)
b

The response function v1g(t) is constructed by superposing


the step response function, Eq.3.54, time shifted and scaled

m
0.13
1
e

us (t )
sec

(3.65)

The response function v1g(t), Eq.3.65, is plotted in Fig.3.68.

3.10.1Example of Second-Order Pulse


Response of an Energized System
Formulate a transfer function for the force source-springmass-damper system equation for the velocity of the mass,
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


the parameter values b = 500 N sec/m, K = 5, 000 N/m, and
M = 400 kg ,
2

1 dF (t ) d v1g b dv1g K
(3.34)
=
+
+ v1g
M dt
M dt
M
dt 2
This system equation allows us to clarify a confusing aspect
of Laplace transformations for the solution of differential
equations. The two initial conditions required to solve a second-order differential equation are the values of the output
variable and its first derivative. In Sect.3.4.2, we established
that if the spring-mass-damper system was de-energized
prior to the application of a step input of force, F (t ) = F0 us (t ),
the initial conditions were the following:

( )

v1g 0+ = 0

and

( )=

dv1g 0
dt

1 dF (t )

=L
M dt

1
L dF (t ) = L
M
dt

150 N us (t 5) + 100 N us (t 10)

2
b dv1g
d v1g
K

2 +L
+ L v1g
M

dt
M dt
2
dv1g K
d v1g b
2 + L
+ L v1g
dt M
dt M

{ }

1
b
K
sF ( s ) = s 2V1g ( s ) +
sV1g ( s ) + V1g ( s )
M
M
M
K
b
1

s + V1g ( s )
sF ( s ) = s 2 +

M
M
M
1
s
M
=
K
b
F (s)
s+
s2 +
M
M

V1g ( s )

1
F0
M

Notice that the initial condition of the first derivative of velocity with respect to time is non-zero, though the system
was de-energized at time t = 0 . How is a non-zero initial
condition to be incorporated into a transfer function method
solution? Must we include the initial condition terms when
performing the Laplace transformation to create the transfer function? No. If the system is initially de-energized, the
initial condition terms are neglected when using transfer
functions. The Laplace transform pair incorporates the contribution of the non-zero initial condition. Again, the system
must be initially de-energized to neglect the initial conditions when using transfer functions. Conveniently, superposition presumes that the system is de-energized before the
first input is applied. We will assume that the previous step
input of 25N was applied at time ta = 6 .
The procedure is as follows:
1. Construct the input function from superposed time shifted
and scaled Heaviside step inputs.
2. Form the transfer function from the system equation.
3. Multiply the transfer function by the transform of a unit step,
to form the Laplace transform of the unit step response.
4. Identify the correct Laplace transform pair. Perform the
algebra needed to create an exact match with the transform pair to find the time-domain unit step response.
5. Construct the pulse response by time shifting and scaling
the unit step response with the same time shifts and scaling factors used to construct the input function.
Construct the input function. Superpose scaled and time
shifted Heaviside step functions. The amplitude of each step
input is the change in the input function at the time that step
function transitions,
F (t ) = 25 N us (t + 9.6) + 75 N us (t )

Form the transfer function

1
s
V1g ( s )
400 kg
=
N sec
N
F (s)
500
5, 000
m s+
m
s2 +
400 kg
400 kg
V1g ( s )
F (s)

0.0025s
s + 1.25s + 12.5
2

The denominator of the transfer function is the characteristic function of the system. Set the characteristic function
equal to zero to form the characteristic equation and solve
it for the systems eigenvalues. If the eigenvalues are real,
then the unit step response is overdamped or non-oscillatory. If the eigenvalues are complex conjugates, then the unit
step response is underdamped or oscillatory:
1.25 1.252 4 12.5
2
s1 , s2 = 0.625 j 3.48.

s 2 + 1.25s + 12.5 = 0 s1 , s2 =

The eigenvalues are complex conjugates. Therefore, the system is underdamped.


The largest time constant in the system is also the systems only time constant. It is the time constant of the decay
envelop of the decaying oscillation. We chose to apply the
preexisting step input six times constants before t = 0, at
ta = 6 .

max = =

1
= 1.6 sec ta = 9.6
0.625

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

164
Fig. 3.69a Force input, F(t).
b Step inputs of force scaled
in amplitude and shifted in time
which superpose to create
the force input, F(t)

100 N u s (t-10)
100

100

75 N us (t)

F(t), N

F(t), N

50

50

10

-6

15

t, sec

25 N us (t+6)

-50

10

15

t, sec

-50

-100

-150 N us (t-5)

-150

The input function shown in Fig.3.69b is


F (t ) = 25 N us (t + 9.6) + 75 N us (t )

150 N us (t 5) + 100 N us (t 10)

Perform the algebra needed to create an exact match with the


form of the transform pair. Multiply the response signal by the
unity ratio, 12.5/12.5. Factor in 12.5 and factor out 0.0025,
V1g ( s ) =

Multiply the transfer function by the transform of a unit


step. The result will be the Laplace transform of the unit step
response,

L {F

u .s.

(t )} =L {1N us (t )} Fu .s. ( s ) = 1NL {us (t )} =


Fu .s. ( s )

V1g ( s ) =

1N
s

u .s.

1
0.0025s

= 2

F ( s ) s s + 1.25s + 12.5

u .s.

0.0025 s

s s 2 + 1.25s + 12.5

V1g ( s ) =
u .s.

0.0025
s 2 + 1.25s + 12.5

e w n t sin w n 1 2 t

w n2
F (s) = 2
s + 2w n s + w n2

Calculate the damping ratio and ideal undamped natural


frequency n by equating the coefficients of like powers of
the Laplace variable, s, in the characteristic functions of the
response signal and the Laplace transform pair,
s 2 + 2w n s + w n2 = s 2 + 1.25s + 12.5

Identify the correct Laplace transform pair to perform the


inverse transformation. The eigenvalues are complex conjugates, so the correct Laplace transform pair is written in
terms of the damping ratio, , and the ideal undamped natural
frequency, n. The correct transform pair is the following:

wn

12.5
0.0025

2
12.5 s + 1.25s + 12.5

12.5

V1g ( s ) = 0.0002 2
s + 1.25s + 12.5
u .s.

V1g ( s )

V1g ( s ) =

f (t ) =

u .s.

12.5
0.0025

2
12.5 s + 1.25s + 12.5

2w n = 1.25

w n = 12.5 = 3.54

and

w n2 = 12.5

rad
1.25
1.25
and =
=
= 0.177
sec
2w n 2 3.54

V 1g ( s ) = L
u .s.

12.5

0.0002 2

1.25
12.5
s
s
+
+

12.5

v1g (t ) = 0.0002L 1 2

u .s
s + 1.25s + 12.5

3.54
e (0.177)(3.54)t sin 3.54 1 0.177 2 t
v1g (t ) = 0.0002
u .s.
1 0.177 2

3.10 Initial Condition of Energized Systems

165

0.6

0.4

v1g(t) 0.2
mm
___
0
sec
-0.2

v1g(t)

cm
___
sec

-3
-6

-0.4
0

t, sec

10

Fig. 3.70 Unit step response of the mass velocity, v1g(t), to a step input
of 1N

-9
-10

-5

10

t, sec

15

20

Fig. 3.71 Response of the mass velocity, v1g(t), in the spring-massdamper 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 )
u .s.

Construct the pulse response by time shifting and scaling


the unit step response function with the same time shifting
and scaling used to construct the input function, F(t),
F (t ) = 25 N us (t + 9.6) + 75 N us (t )

150 N us (t 5) + 100 N us (t 10)

accelerate the mass, FM, and (ii) the force acting through the
dashpot, Fb. The parameter values are b = 500 N sec / m and
M = 200 kg.
Energetic Equations
Continuity or Node Eq : F (t ) Fb FM = 0
Compatibility or Path Eq : v1g = v1g

The response function is given below and plotted in


Fig.3.71,

Element Eqs : Fb = bv1g

FM = M

v1g (t ) = 0.000719e 0.627 (t + 9.6) sin (3.48(t + 9.6)) 25N us (t + 9.6)


+ 0.000719e 0.627 t sin (3.48t ) 75 us (t )

Energy Eqs : E sys = E M

EM =

0.000719e 0.627(t 5) sin (3.48(t 5))150 us (t 5)

+ 0.000719e 0.627(t 10) sin (3.48(t 10))100 us (t 10)

dt

1
Mv12g
2

Reduction i: Input F(t), Output FM


F (t ) Fb FM = 0

F (t ) = Fb + FM

F (t ) = bv1g + FM

3.10.2 Initial Value Method


Superposition is the most efficient method to find the response of a system energized prior to the application of
an input. However, it is possible to view the response to
a step change in the input as creating a new initial value
problem. Thus, we may also create a piecewise continuous response.
We will illustrate the technique with the use of the force
source-mass-damping system of Fig.3.2, but with the damping represented as the alternative energetic schematic symbol
of a dashpot, as given in Fig.3.72. The energetic equations
are reproduced below from Sect.3.2.3.9 for reference. We
will choose as the output variables, (i), the force acting to

dv1g

Differentiate with respect to time to eliminate v1g:


dF (t )
dt

=b

dv1g
dt

dF (t ) b
dFM
dF

=
FM + M
dt
dt
M
dt

This is the system equation. Reorder terms and clear the


zero-order output variable term of its coefficient, in order to
express the system equation in time constant form:
M dF (t ) M dFM
=
+ FM
b dt
b dt

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

166
Fig. 3.72a Translational
mechanical mass-damper system
driven by a force source.
b Linear graph of the massdamper system driven by a
force source

x,v

F(t)

F(t)

g
Check the consistency of the units of the system equation, in
terms of the power variables and time:
Fb = bv1g

F(t), N

F0+

[b] = v

t = 0-

and
FM = M

dv1g
dt

t = 0+
0

Ft
[ M ] = v

M dF (t ) M dFM

=
+ [ FM ]
b dt b dt
M F M F
b t = b t + [ F ]
v F t F v F t F

=
+ [F ]
F v t F v t

Previous
step input

F (t ) = Fb + M

t, sec

F0- = -400 N

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

Check the consistency of the units in terms of the power variables and time:
F
Fb = bv1g [b ] =
v

[F ] = [F ] + [F ]
and

FM = M

The units are consistent.


Reduction ii: Input F(t), Output Fb
F (t ) Fb FM = 0

600 N us(t)

F (t ) = Fb + FM
dv1g
dt

Use the element equation for the damper, Fb = bv1g , to eliminate v1g:
d Fb
M dFb
F (t ) = Fb + M
F (t ) = Fb +
dt b
b dt
This is the system equation. Express it in time constant form
by reordering the output variable terms in decreasing order
of differentiation:
M dFb
F (t ) =
+ Fb
(3.66)
b dt

dv1g
dt

Ft
[M ] =
v

M dFb
M F
F (t ) =
+ [ Fb ] F = b t + [ F ]

b dt
F t v F
+ [F ]
v F t

[F ] =

[F ] = [F ] + [F ]

The units check.


We will now determine the responses of FM and Fb to
the input function, F(t), shown in Fig.3.73. The system is
in steady-state under a step input of force of 400N, when
the input force undergoes a step increase of 1,000N at time
t = 0.
The initial value method views the previous input of
400N as establishing the initial value of the output variable at t = 0+ .

3.10 Initial Condition of Energized Systems

167

Response FM(t) We begin with the response of the force,


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
the instant after a step input is applied, we must first establish 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
the velocity of the mass, v1g,

E sys = E M and E M =

1
Mv12g v1g E storage variable
2

Because the value of an energy storage variable cannot


change instantaneously without infinite power, once we have
established the state variables value at t = 0 , we also know
its value at time t = 0+ ,

E M (0

) = E (0 )
M

1
1
Mv12g 0 = Mv12g 0+
2
2

( )

( )

( )

( )

v1g 0 = v1g 0+

We then use the state variables and input variables values


at t = 0+ in the energetic equations to algebraically determine the value of any power variable in the system at time
t = 0+ .
To determine FM (0+ ) , we must first determine the
value of the state variable, v1g, at time t = 0 in response
to step input, F0 us (t + ta ) = 400 N. We do not know ta, but
we do know that ta occurred earlier enough, that the system reached steady-state in response, by time t = 0 . The
steady-state response of all power variables in the system
to a step input is to reach a constant value. If the state variable, v1g, has reached a constant value, then its derivative
equals zero. Hence, the force acting to accelerate the mass
in steady-state is zero,

( )

v1g ( ss ) = v1g 0 = constant

( )

FM 0 = M

( )

dv1g 0
dt

Steady-State

t = 0+

FM(0+)

~5

t = 0-

time

System in steady-state
under previous step input
Fig. 3.74 Response of the force, FM , acting to accelerate the mass
in the force source-mass-damper system to the input force, plotted in
Fig.3.73

step input is applied to the system, as given in Fig.3.74.


The new input initiates a new dynamic response, which will
reach a new steady-state in five time constants. Time t = 0+
is the beginning of the new transient period. All power variables in the system can change immediately in response to
the new input, except the state (energy storage) variables.
They remain unchanged because no time has elapsed between t = 0 and t = 0+. Consequently, there has been no
time for an energy transfer into or out of the energy storage
element. The transfer of a finite amount of energy requires
a finite amount of time.
The state variable of this system is the velocity of the
mass, v1g. We must establish the value of the state variable
at the instant before the step input is applied, v1g (0 ). We
know the input, F (0 ) = 400 N, and we have established
FM (0 ) = 0. We use these values in the energetic equations
to calculate v1g (0 ),
F (t ) Fb FM = 0

( )

( )

( )

F 0 = Fb 0 + FM 0

We arrive at the critical step in the process. It may appear


that we have the initial condition we need, since FM is the
output variable, and we have established FM (0 ). However,
the force acting to accelerate a mass can change instantaneously,

( )

Transient

Use the element equation of the damper to eliminate Fb:

( )=0

FM 0

FM (t)

( )

FM 0 FM 0+
(3.67)
It is essential to recognize that a steady-state condition exists at time t = 0 , but not at t = 0+ , the instant after the new

F (0 ) = bv1g (0 )

F (0 )
= v1g (0 )
b

m
400 N
= 0.80
= v1g (0 )
N sec
sec
500
m
We can now establish the value of the force acting to accelerate the mass at time t = 0+ , FM (0+ ) . We know the value of
the input at time t = 0+ , F (0+ ) = 600 N. We know that the energy storage (state) variable cannot change from time t = 0
to time t = 0+ ,
m
v1g (0 ) = 0.80
= v1g (0+ )
sec

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

168

1,000

Response Fb(t) The response of the force acting through


the damper to the input plotted in Fig.3.71 follows the
same logic. The initial value method uses the same premise as the method of undetermined coefficients, which is
that the power variables in the system have the functional
form of the input when the system has reached steadystate. We found

800

FM (t), N

600
400
200
0

v1g (0 ) = 0.80

0.5

1.0

1.5

t, sec

2.0

3.0

2.5

Fig. 3.75 Response of the force, FM (t ) , Eq.3.68, acting to accelerate


the mass in the force source-mass-damper system to the input force,
plotted in Fig.3.73

We use the values of the input and the state variable at time
t = 0+ in the energetic equations to determine FM (0+ ). Starting with the continuity equation,

( )

( )

( )

F (t ) Fb FM = 0 F 0+ = Fb 0+ + FM 0+

( )

( )

( )

( )

600 N 500

( )

( )

F 0+ bv1g 0+ = FM 0+

F 0+ Fb 0+ = FM 0+

N sec
m
+
0.80 = FM 0
m
sec

( )

dF0 us ( ss )
M dFM ( ss )
=
+ FM ( ss )
b
dt
dt
0
0


FM ( ss ) = 0

where
M
=
b

200 kg
= 0.40 sec
N sec
500
m

Comparing the initial and final values of the output variable,


FM (0+ ) = 0 < FM ( ss ), we see that the step response is a decay
to zero from initial value of FM (0+ ) = 1, 000 N with the form
t

FM (t ) = FM (0+ )e .
The step response is FM(t), Eq.3.68 below, Fig.3.75,


( )

( )

Fb 0 Fb 0+

The velocity drop, in this case, across the damper and the velocity of the mass are equal, so a conceptual error of equating
Fb (0 ) and Fb (0+ ) would have had no consequence. However, this is not true in general since a damper force can change
instantaneously:

The next step is to find the steady-state value of the output variable, FM(ss), and the time constant using the system
equation in a time constant standard form:

Due to the simplicity of this system, the velocity drop


across the damper equals the velocity of the mass. Since
the force acting through the damper is proportional to the
velocity difference across it, the force in the damper cannot
change instantaneously. However, in general, the force acting through a damping element can change instantaneously.
In general,

(3.69)
Pdamper Pb = Fb v1g

( ) = 1, 000 N

FM 0

m
= v1g (0+ )
sec

FM (t ) = 1, 000 N e 0.40

(3.68)

The rate of energy dissipation can change instantaneously,


because it is not an integral quantity. The accumulation of
energy requires time. Power is the flow rate of energy. The
power flow into an energy storage element is the rate at which
that element accumulates energy. Power can change instantaneously in our simple, abstract models.
The beginning of the new transient response is the moment when we need the initial value of the output variable in order to solve the system equation. Knowing the
input variables value at time t = 0+ and the output variables value at time t = 0 is insufficient, in the general
case, since all power variables in a system can have values
at t = 0+ , which differ from those at time t = 0 , except the
energy storage variables. Therefore, we exploit the special case of the steady-state of a step response, in which
all power variables have constant values, and all derivatives of power variables are zero. We calculate the value of
the energy storage variable in that steady-state condition,
which extends up to time t = 0 . Knowing the input variables and the energy storage variables values at any time,
t, is always sufficient to calculate values of every power
variable in a system.

3.11 Solved Problems

169

800

Knowing the values of the input and the state variable at


time t = 0+, we can now determine the initial value of the
output variable, Fb (0+ ):

600
400

( ) = bv (0 )

Fb 0

1g

Fb(t), N

N sec
m

Fb 0+ = 500
0.80

m
sec

( )

200
0
-200
-400

( )

Fb 0+ = 400N

-600

0.5

1.0

Having the initial value of the output variable, Fb (0 ), we


use the system equation in time constant form to find the
steady-state value, in response to the step input applied at
time t = 0 and identify the time constant:
F0 us ( ss ) =

M dFb ( ss )
+ Fb ( ss )
b
dt 0


Fb ( ss ) = F0 = 600 N for t > 0


where

M
=
b

200 kg
= 0.40 sec
N sec
500
m

Comparing the initial and final values of the output variable we see Fb (0+ ) < Fb ( ss ). Hence, the step response is a
growth from a non-zero initial value over the range

Fb = Fb ( ss ) Fb ( 0+ )

Fb = 600 N ( 400 N )

Fb = 1, 000 N
The step response Fb(t) is plotted in Fig.3.76:
t

Fb (t ) = Fb 1 e + Fb 0+

( )

Fb (t ) = 1, 000 N 1 e

t
0.40

400 N

Rather than using the method of undetermined coefficients,


we will find first-order system step responses by finding the
initial and steady-state values of the output variable and time
constant. We will then identify which is the correct step response function of the four possible options. If the system is
energized prior to the step input, we will use superposition to

2.0

2.5

3.0

Fig. 3.76 Response of the force, Fb, acting through the damper in
the force source mass-damper system to the input force plotted in
Fig.3.73

determine the response, rather than the initial value method.


We will find the step responses of second-order systems, by
first calculating the eigenvalues to determine whether the
system is overdamped and non-oscillatory, or underdamped
and oscillatory. Overdamped systems are easily solved with
either the method of undetermined coefficients or that of
transfer functions. Underdamped systems are most easily
solved using transfer functions.

3.11.1Step Responses of Initially De-energized


System
The example system is the velocity source-spring-damper
system of Sect.3.2.4.2. The energetic schematic and linear
graph of that system are shown in Fig.3.77. The systems
energetic equations and the system equations for the force
acting through the spring, FK, Eq.3.31, and the velocity
drop across the damper, v1g, Eq.3.32, are reproduced below
for reference. The parameter values are K = 5, 000 N/m and
b = 2,500 N sec/m.
Energetic Equations
Continuity or Node Eqs: Fsource = FK

FK = Fb

Compatibility or Path Eq: v1g v (t ) = v12 + v2 g


Element Eqs: Fb = bv2 g

3.11 Solved Problems

1.5

t, sec

Energy Eqs: E sys = E K

dFK
= Kv12
dt
F2
EK = K
2K

bv (t ) =



b dFK
+ FK
K dt

b dv (t ) b dv12
=
+ v12
K dt
K dt

(3.31)
(3.32)

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

170

Fig. 3.77a Energetic schematic


of a velocity source spring-damping system. b Linear graph of the
system shown schematically in
Fig. 3.77a

b
v(t)

x,v

v(t)
g

bv (t ) =

10

v(t)
cm
___
sec 5
0

b dFK
+ FK
K dt

bv0 us (t ) =
1

b dFK ( ss )
+ FK ( ss )
K
dt
0

bv0 = FK ( ss )
0

t, sec

10

15

Fig. 3.78 Step input of velocity acting on the velocity source-springdamper system of Fig.3.75. Note that input units are metric, not SI

Determine the spring force, FK, and the velocity difference


across the spring, v12, when the de-energized system is acted
on by the step input velocity of 10cm/sec, as shown in
Fig.3.78.
Centimeter is not an SI unit. Express the step input in SI
by converting from centimeters to meters:
m
v (t ) = v0 us (t ) v (t ) = 0.1
us (t )
sec
Response Function, FK(t) Identifying the correct step response requires the establishment of initial and steady-state
values, as well as the time constant, . The time constant
is found by unit analysis. The system equation is expressed
in time constant form, when the zero-order output variable
term has no coefficient. The coefficient multiplying the first
derivative of the output variable must have units of time,
bv (t ) =

b dFK
+ FK
K dt

b
K

N sec
m = 0.5 sec
N
5, 000
m

2,500

The steady-state value of the output variable is found by imposing the steady-state condition of a step response that all
power variables in the system reach constant values,

The initial value of the output variable, FK (0+ ), is found


from the energetic equations. The system is described as deenergized, prior to the application of the step input,

E sys ( 0 ) = 0 E sys ( 0 ) = E K ( 0 ) = 0
The amount of energy stored in the system, and in any energy storage element within the system, cannot change instantaneously, since it would require an infinite power flow
to create a finite change over the infinitesimal duration from
t = 0 to t = 0+ . Hence,

E K ( 0 ) = 0 = E K ( 0+ ) E K ( 0+ ) =

( )F

FK2 0+

2K

(0 ) = 0
+

We can summarize that systems which are de-energized,


prior to the application of an input at time t = 0 , remain deenergized at t = 0+ , at the instant after the application of the
input. Consequently, the state (energy storage) variables of
those systems equal zero at time t = 0+.
Comparing FK (0+ ) = 0 with bv0 = FK ( ss ), we identify the
first-order step response as a stable exponential growth from
zero of the form, Eq.3.51,
t
t

FK (t ) = FK ( ss ) 1 e FK (t ) = bv0 1 e

N sec
m

0.1 1 e 2t
FK (t ) = 2,500

m
sec

FK (t ) = 250 N 1 e 2t

Response Function, v12(t) The velocity difference across


the spring, v12 (t ), is found using the same method. Find the
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

300

step responses is the correct response. There is only one time


constant in a first-order system. All system equations must
have the same time constant, =K/b, in the example system.
Similarly, first-order systems have only one independent energy storage element and, consequently, one state variable,
which is the force acting through the spring, FK, in the example system. If we were to find otherwise, then one or both
of the analyses would be erroneous.
The time constant, t, and the steady-state value of the output variable are found from the system equation,
b dv (t ) b dv12
=
+ v12
K dt
K dt

b
= 0.5 sec
K

b dv (t ) b dv12
=
+ v12
K dt
K dt
b dv0 us (t ) 1 b dv12 ( ss )
=
+ v12 ( ss )
dt
K
dt
K
=
0

b dv12 ( ss )
+ v12 ( ss )
K
dt
0

0 = v12 ( ss )

The initial value of the output variable is found from the energetic equations. The system is de-energized, prior to the
application of the step input in velocity. Consequently, the
state (energy storage) variable equals zero, at the instant after
the input is applied, FK (0+ ) = 0 . We always know the value
of the input. In this case, v(0+ ) = 0.1m/ sec. Begin with the
compatibility or path equation, as shown:

( )

( )

( )

v (t ) = v12 + v2 g v 0+ = v12 0+ + v2 g 0+

Eliminate the velocity difference across the damper by expressing, v2g in terms of the damper force, Fb,

( )

( )

v 0+ = v12 0+ +

( )

( )

Fb 0

( )

v 0+ = v12 0+

( )

( )

v 0+ = v12 0+ +

( )

FK 0
b

( )

m 2t
e
sec

t, sec

The response functions FK(t) and v12(t) are plotted in


Fig.3.79.

3.11.2Step Responses of Initially Energized


System
We will find the response of a force source-mass-damper
which is energized, when a step change in force is applied
to the system at t = 0. The key to using superposition to derive the response of a system is to formulate the response to
every input assuming that the system is de-energized when
that input is applied, even if it is clearly not. One assumes
that the system is de-energized, because the effect of the nonzero initial conditions will be accounted for, as the response
of the prior input.
We will rework the example of Sect.3.10.2. The force
source-mass-damper system is shown in Fig.3.80a, and
the input is shown in Fig.3.80b. The output variables are
(i) the force acting to accelerate the mass, FM, and (ii) the
force acting through the dashpot, Fb. The energetic equations and the system equations are reproduced below from
Sects.3.2.3.9 and 3.10.2 for reference. The parameter values
are b = 500 Nsec / m and M = 200 kg.
Energetic Equations
Continuity or Node Eq: F (t ) Fb FM = 0
Compatibility or Path Eq: v1g = v1g

m
with v12 ( ss ) = 0, we see that the
sec
response is a decay to zero of the form of Eq.3.50,
t

v12(t)

100

0.02

Element Eqs: Fb = bv1g

FM = M

Energy Eqs: E sys = E M

EM =

Comparing v12 (0+ ) = 0.1

( )

v12(t)
m
__
0.04 sec
0.06

m
0.1
= v12 0+
sec

v12 (t ) = v 0+ e v12 (t ) = 0.1

0.08

200

Fig. 3.79 Step responses, FK(t) and v12(t)

The time constant is the same as found in the previous analysis. 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,

b dv0
K dt

FK(t)
N

0.1

FK(t)




dv1g
dt

1
Mv12g
2

M dF (t ) M dFM
=
+ FM
b dt
b dt
F (t ) =

M dFb
+ Fb
b dt

(3.26)
(3.59)

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

172
Fig. 3.80a Energetic schematic
of a mass-damping system acted
on by a force source. b The
system was in steady-state, prior
to step change of the input from
F (0 ) = 400 N to F (0+ ) = 600 N

F(t), N
F0+

x,v

F(t)

t = 0-

t = 0+

0
Previous
step input

Response Function, FM(t) We will first determine the unit


step response for the output variable, FM, assuming that the
system is initially de-energized. Then we shall scale, time
shift, and sum these responses to correspond to the scaling
and time shifting of Heaviside step functions superposed to
create the input function.
Determine the unit step response assuming that the system is initially de-energized. Express the time constant and
the steady-state value of the step response from the system
equation in time constant form:
M dF (t ) M dFM (t )
=
+ FM (t )
b dt
b
dt

M
b

1
Mv12g 0 = 0
2

( )

The amount of energy stored cannot change instantaneously.


Therefore, it follows that

E M (0 ) = E M (0+ ) =

1
Mv12g (0+ ) = 0 v1g (0+ ) = 0
2

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
output variable at time t = 0+ , F M (0+ ):
u .s.

( )

M d
M dFM ( ss )
+ FM ( ss )
1 N us ( ss ) =
1
b dt
b
dt

E sys ( 0 ) = E M ( 0 ) = 0 E M ( 0 ) =

( )

( )

FM 0+ = F 0+ b v1g 0+

M dFM (t )
M d
1 N u s (t ) =
+ FM (t )
b
dt
b dt

F0 - = -400 N

( )

Determine the steady-state unit step response value of the


output variable, FM. The steady-state response of power variables in any system to a step is to reach a constant value:

M dFM ( ss )
=
+ FM ( ss )
b
dt
0

t, sec

( )

( )

F Fb FM = 0 FM 0+ = F 0+ Fb 0+

200 kg
= 0.40 sec
=
N sec
500
m

M d 1N
b dt

600 N us(t)

FM ( ss ) = 0

Use the algebraic equations of the equation list to determine


the initial value of the output variable, FM (0+ ), assuming the
system is de-energized. The first step is always to establish
the value of the state (energy storage) variable at time t = 0+ ,
the instant after the step input is applied.
If a system is de-energized, then all of its energy storage
(state) variables equal zero. First-order systems have one independent energy storage element. The energy storage element in this system is mass,

( )

FM 0+ = 1N
u .s.

Compare the initial and steady-state values of the output


variable to identify the type of step response:
FM (0+ ) = 1N and FM ( ss ) = 0
The unit step response is a decay to zero of the form

( )

FM (t ) = FM 0+ e
u .s.

( )

FM (t ) = FM 0+ e
u .s

b
t
M

FM (t ) = 1N e 2.5 t
u .s

Construct the input function, F(t), by superposing (summing) scaled and time shifted Heaviside step functions,
F (t ) = F1 (t ) + F2 (t ), Fig.3.81,
 F (t ) = 400 Nus (t + ta ) + 1, 000 Nus (t )

where ta 5
(3.70)

Construct the response function, FM (t ) = FM1 (t ) + FM 2 (t ) ,


by superposing (summing) scaled and time shifted unit step

3.11 Solved Problems

173
F2(t) = 1,000 N us(t)

F(t), N
1,000

1,000

FM(t)

800

Step Input 2

600

t = 0-

ta

t = 0+

F1(t) = -400 N us(t+t a )

t, sec

-400

Fig. 3.81 Input step function, F (t ), created by superposing (summing)


scaled and time shifted heaviside step functions

functions, using the same scaling and time shifting used to


create the input function
FM (t ) = 400us (t + ta )1Ne

b
( t + ta )
M

FM (t ) = 400N us (t + 2) e

2.5(t + ta )

+ 1, 000 N us (t )e

b
t
M

M dFb
+ Fb
b dt

M
= 0.4 sec
b

If we had found a different time constant than that from the


system equation for FM, one or both constants would be erroneous. The steady-state value of the output variables
unit step response is found by the condition that it must be
constant,
M dFb
+ Fb
F (t ) =
b dt
1

Fb(t)
0

M dFb ( ss )
+ Fb ( ss )
b
dt
0

1N = Fb ( ss )

To determine the value of the output variable, Fb, at the instant after the unit step is applied to the system, we must
first determine the value of the state variable at that instant.
This is particularly easy, since we must assume that the system is de-energized to use superposition. If the system is
de-energized, then we know that the state (energy storage)

0.5

1.0

1.5

2.0

t, sec

2.5

3.0

Fig. 3.82 Response functions of the forces acting to accelerate the


mass, FM(t), and acting through the damper, Fb(t)

variables equal zero. State variables cannot change instantaneously, since that would require infinite power. Therefore,

( )

( )

v1g 0 = 0 = v1g 0+

+ 1, 000N us (t ) e 2.5t

Response Function, Fb(t) We will again use superposition


to determine the response function, by applying the preexisting step input at a time ta such that the system has reached
steady-state by time t=0, and then superpose (sum) the responses due to the inputs applied at time t=ta and time t=0
to construct the response function with the same time shifts
and scaling as the input function.
Determine unit step response. The time constant and
the steady-state values of the output variable are determined
from the system equation. The time constant is shown here:

1N us (t ) =

0
-200

Step Input 1

F (t ) =

FM(t), N 400
Fb(t), N 200

We now use the values of the input and the state variable at
time t = 0+ to determine the initial value of the output variable,
Fb (0+ ). As it happens, we only need the value of the state variable, because the force acting through the damper is proportional to the velocity difference between nodes one and ground,

( )

( )

Fb 0+ = b v1g 0+

( )

Fb 0+ = 0

Identify the correct first-order step response from the initial and steady-state values of the output variable, Fb (0+ ) = 0
and Fb ( ss ) = 1 N . The unit step response is one of stable exponential growth,
t

Fb (t ) = F0 1 e

b
t

Fb (t ) = 1N 1 e M

u .s.

Fb (t ) = 1N 1 e 2.5 t
u .s.

Construct the response function scaling and time shifting


the unit step response function, by means of the same scaling
and time shifting used in the input function. We constructed
the input function, F(t), Eq.3.70, for our calculation of the
response, FM. We employ Eq.3.70, but remove the units of
newtons, since the unit step response function carries units,
yielding
b

( t + ta )
Fb (t ) = 1N 1 e M
400 us (t + ta )

b
t

+ 1N 1 e M 1, 000 us (t )

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

174

x,v

v(t)

m
___
v(t), sec

0.8
0.6
0.4
0.2
2

Frictionless rollers
Fig. 3.83 Energetic schematic of a velocity source-damper-massspring system

10

12

14

t, sec

-0.4

Fig. 3.84 The system was in steady-state under a step input of


0.25m/sec, when the pulse input shown was applied at time, t=0+

and, substituting for the time constant,

Fb (t ) = 1N 1 e

2.5 (t + ta )

) 400 u (t + 2)

x,v

+ 1N 1 e 2.5 t 1, 000 us (t )

g v(t)

b
M

3.11.3Second-Order Step Responses and Pulse


Responses
We will use the velocity source-damper-mass-spring system
shown in Fig.3.83 as the example system. The velocity source
is a machine or machine element which moves with a known
(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
graph method, we will derive system equations for (i) the force
acting to accelerate the mass and (ii) the force acting through
the spring.
The response of a system to any arbitrary pulse, such as
the input function, v(t), shown in Fig.3.84, is constructed
using superposition. We will solve the system equations for
two sets of parameters, in order to illustrate solution methods for both the overdamped, non-oscillatory and the underdamped, oscillatory cases.
The key step in drawing a linear graph from an energetic schematic is to locate the nodes of distinct values of
the across variable, Fig.3.85. The across variable in a mechanical system is velocity. The only energetic attribute of
an ideal mass is kinetic energy storage. It cannot deform,
or it would either store strain energy or dissipate mechanical energy as heat through the shear of plastic deformation.
Therefore, a mass is perfectly rigid with a single velocity.
The inertial reference velocity for the mass is ground. The
velocity source must react against something to produce
the force needed to impose the prescribed velocity at its
point of action. Since nothing is shown in the schematic,
one assumes that it reacts against ground. The velocity
source, damper, and spring have nodes at either end, since
their ends can move independently.
Draw the linear graph by first drawing and labeling the
velocity nodes, Fig.3.86. Then add branches between the

Frictionless rollers
Fig. 3.85 Nodes of distinct values of the across variable velocity
shown on the energetic schematic of a velocity source-damper-massspring system

respective nodes to represent energetic elements. Define


the positive direction of the through variable force in each
branch with an arrowhead. Sources increase the across variable in the positive direction of the through variable. In all
other elements, the across variable decreases in the positive
direction of the through variable.
Write one continuity equation less than there are nodes.
There are three nodes. The two ground nodes are of the same
velocity and represent the same node. Write continuity equations for nodes one and two. Omit the ground node. Notice
that force applied by the velocity source equals force acting through the damper, Fb, whereas the force Fb divides between the force, FM, acting to accelerate the mass, and the
force, FK, acting through the spring.
There can be no more useful compatibility equations than
there are holes in the linear graph. There are two holes. In
this system there is only one useful compatibility equation.
The second is either a restatement of the first or a trivial
equation.
Energetic Equations
Continuity or Node Eqs: Fsource = Fb

Fb = FM FK

Compatibility or Path Eqs: v1g v (t ) = v12 + v2 g


v2 g = v2 g

3.11 Solved Problems

175

Fig. 3.86a Nodes of distinct


values of the across variable
velocity. b Linear graph of the
velocity source damper-massspring system

v(t)

Energy Eqs: E sys = E M + E K

dv2 g

dFK
= Kv2 g
dt
dt
F2
1
E M = Mv22g E K = K
2
2K

Element Eqs: Fb = bv12 FM = M

Reduction i: Input v(t), Output FM


v (t ) = v12 + v2 g
dv (t )

1 dFb dv2 g
=
+
dt
b dt
dt
dv (t )
dt

dt

d 2 v (t )
dt 2

1 d
1
=
FM + FK ) +
FM
(
dt
b dt
M

1 dFK 1 dFM
1
FM
+
+
b dt
b dt
M

1 dFM
1
K
v1g +
+
FM
b
b dt
M

dt
2

dv (t )

1 dFM
K dv1g 1 d 2 FM
+
+
b dt
b dt 2
M dt

K
1 d 2 FM
1 dFM
FM +
+
2
bM
b dt
M dt

d 2 v (t )
dt

d 2 v (t )
dt

d 2 FM
b dFM K
FM
+
+
2
M dt
M
dt

F
F = bv [b ] =
v

F=M

dv
F t
[M ] =
dt
v

d 2 v (t ) d 2 FM
= 2
b
2
dt dt

b dFM K

+
+ M FM
M
dt

F v F F v F v F
= 2+
F

+
2
v t t v F t t F t t v
F F F F
t 2 = t 2 + t 2 + t 2

1 d 2 FM
1 dFM
K
+
+
FM
bM
b dt 2
M dt
=

Check the consistency of the units in terms of the power variables and time:

v F b F K
b t 2 = t 2 + M t + M F

This is the system equation, since the only power variables


which appear are the input and output variables. Rearrange
the output variable terms in the decreasing order of differentiation. Clear the coefficient from the second-order term to
express the system equation in standard form:

(3.71)

F
dF
= Kv [ K ] =
dt
t v

1
Fb + v2 g
b

dv (t )
d 2 v (t )

v (t ) =

The units are consistent.


Reduction ii: Input v(t), Output FK
v (t ) = v12 + v2 g

v (t ) =

1
1 dFK
Fb +
b
K dt

v (t ) =

1
1 dFK
FM + FK ) +
(
b
K dt

v (t ) =

1
1
1 dFK
FM + FK +
b
b
K dt

v (t ) =
v(t ) =

1 dFK
M dv2 g 1
+ FK +
b dt
b
K dt

1 dFK
M d 2 FK 1
+ FK +
2
bK dt
K dt
b

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

176

m
v(t), ___
sec

duration before t=0 of five times the maximum constant of


the system, ta = 5 max,

0.6
0.4
0.2

t, sec

-5max

v(t ) = 0.2

10 12 14 16

-0.4

m
m
m
us (t + 5 max ) + 0.4 us (t ) + 0.6 us (t 2)
sec
sec
sec
m
m
1.2 us (t 5) + 0.4 us (t 10)
sec
sec

-0.6

(3.73)

-0.8
-1.0
-1.2

Fig. 3.87 The input function, v(t), is constructed of five scaled and
time shifted heaviside step functions

This is the system equation for the force acting through


the spring. Express it in standard form by rearranging the
derivatives in decreasing order. Remove the coefficient from
the second-order output variable term:
v(t ) =

M d 2 FK 1 dFK 1
+
+ FK
K dt
b
bK dt 2

d 2 FK
(3.72)
bK
b dFK K
v(t ) =
+
+
FK
M
M dt
M
dt 2
Check the units of the system equation for consistency:
F = bv

[b] = v

F=M

dF
= Kv
dt

dv

dt

3.11.3.1Pulse Response of the Overdamped


System
The parameter values b = 600 N sec / m, M = 200 kg, and
K = 400 N / m yield an overdamped system.
Pulse response of the force accelerating the mass, FM(t)
Create the transfer function. Perform the Laplace transformation on the system equation, Eq.3.71, neglecting the
initial condition terms:

2
d 2F

b dFM K
d v (t )
b
FM
= L 2M +
+

2
M
dt
M
dt
dt

bs 2V ( s ) = s 2 FM ( s ) +

Ft
v

[ M ] =

F b F K
v = 2 +
F
+
t M t M

F F v
F F v F F v

v = 2+
F

+
v t v F t t v F t t t v F t

Form the ratio of the output variable over the input variable:
Output ( s ) FM ( s )
bs 2
=
=
(3.74)
b
K
Input ( s )
V (s)
s2 +
s+
M
M
Calculate the eigenvalues to establish whether the system is
underdamped or overdamped. The denominator of the transfer function is the characteristic function. Set it equal to zero
to form the characteristic equation:

F F F F
t 2 = t 2 + t 2 + t 2
The units are consistent.
Construct the input function by scaling and time shifting Heaviside step function. The input, v(t), Fig.3.84, is the
superposition of five step functions, Fig.3.87. The system is
described as in steady-state at time t=0 in response to a step
input of 2.5 m/sec , applied previously. The time at which
this input was applied to the system is unknown, but the fact
the system has reached steady-state provides the minimum

b
K
sFM ( s ) +
FM ( s )
M
M

b
K

bs 2V ( s ) = s 2 +
s + FM ( s )

M
M

[K ] = t v

2
bK
d FK b dFK K

v
t
=
+
(
)

2
+ M FK
M

M
dt
dt

bK
M

We will now calculate the responses of the system to the


pulse input, beginning with the overdamped case. The pulse
responses are constructed by superposing scaled and time
shifted unit step responses. We will determine the step responses via transfer functions.

s2 +
s2 +

s1 , s2 =

b
K
=0
s+
M
M

600
400
s+
= 0 s 2 + 3s + 2 = 0
200
200

(3)2 4 2
2

s1 = 1 and s2 = 2

Real eigenvalues indicate the system is overdamped or nonoscillatory.

3.11 Solved Problems

177

600

The maximum time constant of the system equals the absolute value of the smallest real component of the eigenvalues of the system,

max =

min

max =

1
= 1sec
1

400

FM(t), N
u.s.

Find the Laplace transformation of the input using the


Laplace transform pairs of Table2.3. The input is a unit step,
v(t )=us (t ) . The relevant Laplace transform pair is
us (t ) = 0 for t < 0
u (t ) = 1 for t > 0
s

F s = 1
( ) s

0
-100
-1

L {v (t )} = L {u (t )}

V (s) =

FM ( s ) 1
=
V (s) s

FM ( s ) =

1
s

bs 2
b
K
s2 +
s+
M
M

bs 2
b
K

s s2 +
s+

M
M

bs
600 s
FM ( s ) = 2
FM ( s ) =
b
K
s
+
3s + 2
s2 +
s+
M
M
Perform the inverse Laplace transformation by using the
Laplace transform pair to return to the time-domain. Examine Table2.3 to seek a Laplace transform pair with the same
form. The only signals with second-order denominators in
polynomial form correspond to oscillatory time-domain responses. The overdamped second-order pairs are written in
factored form. The relevant Laplace transform pair for the
inverse transformation unit step response is below,
1

bt
at
f (t ) = b a be ae

s
F (s) =
s
a
+
( )( s + b)

Factor the denominator of FM(s):


s2 +

K
b
s+
= s 2 + 3s + 2 ( s + a ) ( s + b ) = ( s + 1) ( s + 2)
M
M

Multiply the transfer function by the Laplace transformation of the input to yield the Laplace transformation of
the output:
V (s)

t, sec

Fig. 3.88 The unit step response of the force acting to accelerate the
mass, FM(t)

The Laplace transform of the unit step input is


v (t ) = us (t )

200

Take care not to confuse the meaning of variables. The b in


the Laplace transform table is the opposite of an eigenvalue.
The b in FM(s), the Laplace transform of the response function, is the damping coefficient of the system.

FM ( s ) =
u .s.

600 s

( s + 1)( s + 2)

FM ( s ) = L
u .s.

FM (t ) = 600L
u .s

600 s

+
1
+
2
s
s
)( )
(

( s + 1)( s + 2)

FM (t ) =

600N
be bt ae at
ba

FM (t ) =

600 N
2e 2t 1e 1t
2 1

u .s.

u .s.

The unit step response is below and plotted in Fig.3.88:


FM (t ) = 1, 200Ne 2t 600Ne t
(3.75)
u .s.

The unit step response of the force acting to accelerate the


mass, Eq.3.75, undershoots its steady-state value of zero. Is
the system oscillatory, even though the eigenvalues are real?
No, the system is not oscillatory. The power variable plotted, FM, is the force acting to accelerate the mass. It reverses
signs, because the initially de-energized (at rest) mass must
move to compress the spring and damper to force transfer
to it. What we see in the plot is the mass accelerating and
decelerating, which is necessary, if the mass begins and ends
at rest. If the velocity of a mass changes sign during a step
response, then the system is underdamped and oscillatory.
Use superposition to construct the pulse response function FM(t), Fig.3.89. Scale and time shift the unit step re-

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

178

300

bK
Output ( s ) FK ( s )
M
(3.77)
=
=
b
K
Input ( s )
V (s)
2
s + s+
M
M

200
100

FM(t), N 0

Substituting in the parameter values yields, as below,

-100

FK ( s )
=
V (s)

-200
-300
-6

-3

t, sec

12

15

Fig. 3.89 The response of the force acting to accelerate the mass,
FM(t), Eq.3.76, to the pulse input, Eq.3.73, shown in Fig.3.84

sponse function, Eq.3.75, using the same scaling and time


shift used to construct the input function, v(t), Eq.3.73, to
create the pulse response function Eq. 3.76:
m
m
m
v (t ) = 0.2 us (t + 5 max ) + 0.4 us (t ) + 0.6 us (t 2)
sec
sec
sec
m
m
1.2 us (t 5) + 0.4 us (t 10)
sec
sec

FM (t ) = 0.2 1, 200Ne 2(t + 5) 600Ne (t + 5) us (t + 5)

+ 0.4 1, 200Ne 2t 600Ne t us (t )

(
1.2(1, 200Ne (
+ 0.4(1, 200Ne (

2 t 5

)
)
600Ne ( ) )u (t 5)
)
600Ne ( ) )u (t 10)

2 t 10

t 5

t 10

(3.76)

Pulse response of the force acting through the spring, FK(t)


Create the transfer function by performing the Laplace
transformation on the system equation, Eq.3.66, neglecting
the initial condition terms:

Multiply the transfer function by unit step to yield the


Laplace transform of the unit step response:
v(t )=us (t )
V (s)

d 2 FK

b dFK K
bK

=
+
v
t
FK
(
)
L

2 +
M dt
M
M

dt

K
b
bK
V ( s ) = s 2 FK ( s )+ sFK ( s ) +
FK ( s )
M
M
M
bK
b
K

V ( s ) = s 2 + s + FK ( s )

M
M
M
The transfer function is the ratio of the output variable over
the input variable,

V ( s )=

1
s

FK ( s ) 1 1, 200
1, 200
=
FK ( s ) =
2
2
V ( s ) s s + 3s + 2
u .s.
s s + 3s + 2

FK

(s) =

1, 200
2

s s + 3s + 2

FK

u .s.

(s) =

1, 200
s ( s + 1)( s + 2)

The relevant Laplace transform pair is

1
1
at
bt
f (t ) = ab 1 + a b be ae

1
F (s) =

s ( s + a )( s + b )

FK ( s ) = L
u .s.

1, 200

s ( s + 1)( s + 2)

1
FK (t ) = 1, 200L

u .s.
s ( s + 1)( s + 2)
1

FK (t ) =
u .s.

2
bK
L {v(t )} = L d F2K + b L dFK + K L { FK }
M
dt M
dt M

L {v(t )} = L {u (t )}

Perform the inverse Laplace transformation using the Laplace transform pair.
Factor the denominator to match the form of overdamped
signals:

u .s.

+ 0.6 1, 200Ne 2(t 2) 600Ne (t 2) us (t 2)

600 400
F (s)
1, 200
200
K
= 2
600
400
V ( s ) s + 3s + 2
s2 +
s+
200
200

1, 200
1

1+
2e 1t 1e 2t

1 2 1 2

( (

FK (t ) = 600 1 2e t e 2t
u .s.

))

The unit step response of the spring force, FK(t), Eq.3.78


below, is plotted in Fig.3.90,

FK (t ) = 600 1 2e t + e 2t
(3.78)
u .s.

Note that the unit step response of the spring force, Fig.3.90,
does not overshoot its steady-state value.

3.11 Solved Problems

179

600

600

FK (t), N

400

400

FK(t), N

u.s.

200

200
0

-200

t, sec

Fig. 3.90 The unit step response of the spring force, FK(t), Eq.3.78

Use superposition to construct the pulse response


function. Use the same scaling for the time shifting, as was
used to create the input function v(t), Eq.3.73, Fig.3.91:
v (t ) = 0.2

m
m
m
us (t + 5 max ) + 0.4
us (t ) + 0.6
us (t 2)
sec
sec
sec
1.2

m
m
us (t 5) + 0.4
us (t 10)
sec
sec

( (

))

FK (t ) = 0.2 600 1 2e (t + 5) + e 2(t + 5) us (t + 5)

( (
+ 0.6 ( 600 (1 2e (
1.2 ( 600 (1 2e (
+ 0.4 ( 600 (1 2e (

))

+ 0.4 600 1 2e t + e 2t us (t )

))
)
+ e ( ) )) u ( t 5 )
)
+ e ( ) )) u (t 10)

t 2)

t 5

t 10

+ e 2(t 2) us (t 2)
2 t 5

2 t 10

(3.79)

3.11.3.2Pulse Response of the Underdamped


System
The parameter values b = 400 N sec/m , M = 200 kg, and
K = 2, 000 N/m yield an underdamped system.
Pulse response of the force accelerating the mass, FM(t)
The transfer function for the force acting to accelerate
the mass, FM(t), is Eq.3.74, reproduced below,
Output ( s ) FM ( s )
bs 2
(3.74)
=
=
b
K
Input ( s )
V (s)
s2 +
s+
M
M
Calculate the eigenvalues. The denominator of the transfer
function is the characteristic function,
s2 +
s2 +

b
K
=0
s+
M
M

400
2, 000
s+
=0
200
200

s 2 + 2 s + 10 = 0

-400

-6

-3

t, sec

12

15

18

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

s1 , s2 =

( 2)2 4 10
2

s1 , s2 = 1 j 3

The eigenvalues are complex conjugates, indicating that the


system is underdamped and oscillatory. The frequency of oscillation is magnitude of the imagery component, 3rad/sec.
The real component is the coefficient of the exponent of the
decay envelope.
The time constant of the system equals the absolute value
of the real component of the eigenvalues of the system,
1
1
=
=
= 1sec.
w
1
The unit step response is the product of the Laplace
transform of a unit step and the transfer function. The Laplace transform of the unit step input is
v(t )=us (t )

L {v(t )} = L {u (t )}
s

Input ( s )
V (s)

FM ( s ) =

V ( s )=

1
s

Output ( s )
= Output ( s )
Input ( s )

FM ( s ) 1
bs 2
=
V (s)
s s2 + b s + K
M
M

bs
400 s
FM ( s ) = 2
b
K
s + 2 s + 10
s2 +
s+
M
M

The relevant Laplace transform pair is written in terms of the


damping ratio, , and the ideal undamped natural frequency,
n, because the response is underdamped and oscillatory,

e w n t sin w n 1 2 t
f (t ) =
2

1 2
wd

1
1
where
tan
tan

=
=

s
F (s) =

s 2 + 2w n s + w n2

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

180

FM(t), N

300

300

200

200
100

100

FM(t), N 0

-100

-100

-200

-200
0

t, sec

Fig. 3.92 The unit step response of the force acting to accelerate the
mass, FM(t)

F M ( s ) = L
.
.
u
s

F M (t ) = 400L

400 s
2

s + 2 s + 10

u .s.

w n = 10 = 3.16
=

wn

rad
sec

s1 10 = w n2
s 0 2 = 2w n

2 = 2w n

1
= 0.316
3.16

1 2
where = tan 1


1
e

w n t

=e

12

15

F M (t ) = 422e t sin (3t 0.69 rad )


(3.80)
u .s.

The pulse response, FM (t), Eq. 3.81, Fig.3.93, is created by


scaling, time shifting, and superposing the unit step response
with the same scaling factors and time shifts as the input
function v(t), Eq.3.73, below,
v (t ) = 0.2

m
m
m
us (t + 5 max ) + 0.4 us (t ) + 0.6 us (t 2)
sec
sec
sec
m
m
1.2 us (t 5) + 0.4 us (t 10)
sec
sec

+ 0.4 422e sin (3t 0.69 rad ) us (t )


t

(
1.2( 422e (
+ 0.4( 422e (
+ 0.6 422e

(t 2 )

t 5)
t 10

)
sin (3 (t 5) 0.69 rad ))u (t 5)
)
sin (3 (t 10) 0.69 rad ))u (t 10)
sin (3 (t 2) 0.69 rad ) us (t 2)
s

(3.81)

400
1 ( 0.316)
(0.316)(3.16)t

Fig. 3.93 The pulse response of the force acting to accelerate the mass,
FM(t), Eq.3.81

1
e w n t sin w n 1 2 t
FM (t ) = 400
2
u .s.
1

400

t, sec

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


mass, FM(t), is

The unit step response of FM(t), Eq.3.80, is plotted in Fig.3.92,

400 s
400 s
=
s 2 + 2 s + 10 s 2 + 2w n s + w n2

1 ( 0.316)2
1 2
1
= 0.69 rad
= tan
= tan
0.316

The damping ratio, , and the ideal natural frequency, n, are


calculated by equating the coefficients of like powers of s in
the denominator,

s 2 + 2 s + 10 = 2w n s + w n2

-3

s
s
2
10
+
+

-300
-6

=e

= 422
t

w n 1 2 t = 3.16 1 ( 0.316) t = 3t
2

Pulse response of the force through the spring, FK(t)


The transfer function for the force acting to accelerate
the mass, FK(t), is Eq.3.77, from above,
bK
Output ( s ) FK ( s )
M
(3.77)
=
=
b
K
Input ( s )
V (s)
2
s + s+
M
M

3.11 Solved Problems

181

600

Evaluate the transfer function using the parameter values for


an underdamped system, b = 400 N sec/m , M = 200 kg, and
K = 2, 000 N/m:
400 2, 000
FK ( s )
F (s)
4, 000
200
=
K
=
400
2, 000
V ( s)
V ( s ) s 2 +2 s + 10
2
s +
s+
200
200

500

FK (t), N
u.s.

400
300
200
100

Multiply the transfer function by the Laplace transform of a unit


step to yield the Laplace transform of the unit step response:
V (s)

FK ( s ) 1 4, 000
4, 000
=
FK ( s ) =
V ( s ) s s 2 +2s + 10
s ( s 2 +2s + 10)

FK ( s ) =
u .s.

FK ( s ) = 400
u .s.

10
s ( s +2 s + 10)
2

The damping ratio and the ideal, undamped natural frequency for this system and parameters, as calculated above, are

= 0.316 and w n = 3.16

rad
sec

t, sec

12

15

200

FK(t), N 0
-200
-400
-6

10
4,000
4,000
10
FK ( s ) =
10 s ( s 2 +2 s + 10)
10 s ( s 2 +2 s + 10)
u .s.

400

There must be an exact match with the Laplace-domain


function of the transform pair. Multiply the response function by the unity ratio, 10/10:

Fig. 3.94 The unit step of the spring force, FK(t), Eq.3.82

The relevant Laplace transform pair from Table2.3 is the


following:

1
e w n t sin w n 1 2 t +
f (t ) = 1
2

1 2
wd

1
1

where
tan
tan
=
=

w n2
F (s) =

s s 2 + 2w n s + w n2

-3

t, sec

Fig. 3.95 The response of the spring force, FK(t), to the input function,
v(t), Eq.3.66

The unit step response of the spring force FK(t) is Eq.3.82,


below, and is plotted in Fig.3.94,
FK (t ) = 400 422e t sin (3t + 1.25 rad )
(3.82)
u .s.

The unit step response is scaled, time shifted, and superposed to create the pulse response, FK(t), Fig.3.95, for the
input function, v(t), Eq.3.73,
v (t ) = 0.2

Perform the inverse Laplace transformation using the transform pair:

m
m
m
us (t + 5 max ) + 0.4 us (t ) + 0.6 us (t 2)
sec
sec
sec
m
m
1.2 us (t 5) + 0.4 us (t 10)
sec
sec

1
2
e (0.316)(3.16)t sin 3.16 1 ( 0.316) t + 1.25 rad
FK (t ) = 400 1
2
u .s.

1 ( 0.316)

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

182

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
Linear graphs are circuit-like representations of energetic
systems. Draw a linear graph by drawing and identifying
the velocity nodes. Then add the elements between the
nodes. The crucial step is to identify all nodes of the across
variable on the schematic of the system. The symbol for a
source is a circle with the source variable in it. The other
branches are lines with an arrowhead, indicating the positive direction of the through variable (in this case, force)
and the positive direction of the drop, or decrease in the
across variable, velocity. The lower half of the branch M is
a dashed line to indicate that no reaction force acts on the
mass from the ground.
Write all equations which can be stated for the energetic
model, represented by the linear graph before beginning any
algebra. Organize the equations for easy reference. The equations which describe an energetic system are categorized
as continuity (node), compatibility (loop or path), element
(constitutive), and energy. They will be used to derive the
differential system equation and then to establish the initial
conditions needed to solve the system equation.
The compatibility and continuity equations depend solely
on how elements are connected to form a network. All energetic systems have a variable which sums at a point or node.
In mechanical systems, the through variables, forces and
torques, sum at a node to yield equilibrium equations. We
write one node equation less than there are nodes in the system. It is best to omit the ground node from the node equations. Similarly, all energetic systems have a variable which
changes between nodes. In mechanical systems, the across
variable is velocity. The differences between the values at
nodes must be consistent and yield the same sum along two
different paths between node A and node B, or zero when
summed around a loop back to the starting node. We can
write no more, and often fewer, independent loop or path
equations than there are holes in the linear graph.
The element or constitutive equations are the model of a
single energetic attribute. The element equations, which relate the power variables of individual elements and energy
equations, which state the amount of energy stored in the
system, must be written with the nodes and assumed positive direction of the through variables defined on the linear
graph. We must use linear element equations to yield linear
differential equations. The energy equations are statements

of which elements in the system store energy, and how the


energy stored in an element is calculated.
The set of energetic equations is reduced using elimination by substitution to derive the differential system equation, which relates the input variable to an output variable
of interest.
Solution of differential system equations. We have two
complementary ways of solving the system equation for an
input function constructed by superposing scaled and time
shifted Heaviside step functions. The method of superposing scaled and time shifted unit step responses is much more
efficient than the initial condition method. The initial condition method requires a new solution for each segment of the
input pulse, whereas superposition method does not.
Initial condition method. View the previous inputs as establishing the initial value of the output variable for the next
input. The response of the system to the input establishes
the value of the output variable at time t = 0 . Knowing the
values of the input and output variables at time t = 0 , and
also that the steady-state values of all variables in the system
must be constants, provide sufficient information to establish the value of the energy storage variable at time t = 0 .
The Heaviside step input transitions instantaneously from
zero to one when its argument becomes positive. Although
we model the inputs power variable as capable of an instantaneous change, an energy storage (state) variables value
cannot change instantaneously. Power must flow into or out
of an energy storage element for the state variables value
to change. Flow takes time. Only infinite flow rates would
allow finite changes over zero time. Infinite power flows are
impossible. Therefore, the energy (state) variables value at
times t = 0 and t = 0+ are equal. Important. The output variable of the system equation may, or may not, be the energy
storage (state) variable. In any case, the only variables whose
values are known at time t = 0+ , at the instant after a step
input has transitioned, are the input variable and the state
variables. If the output variable is not the state variable, then
it must be determined by using the input and state variables
values at t = 0+ in the energetic equations.
Superposition method. Sum the responses due to the
previous input and the input applied at time t=0. First,
create the input function as the sum of step functions, except
that the step input applied at an unknown negative time must
be represented as a constant, since we do not know when its
transition occurred, other than it was long enough ago, that

Problems

the systems response to it reached steady-state. We solve


for the response of a unit step of the input variable, as if that
input acts on a de-energized system, even when is clearly not
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
inputs, which are superposed to create the input pulse;
2. Time shift the scaled unit step responses, by the same
time shifts used to create the pulse input;
3. Multiply each term of the response function by a Heaviside unit step function with the same time shift, in order to
zero-out that term, until its corresponding input step has
acted on the system; and
4. Sum (superpose) the response functions which are scaled
and shifted in time as the individual input step function,
which creates that response function.
When the system is de-energized, before a step input is applied to the system, we know that the values of the energy
storage (state) variables at time t = 0 are zero. We also
know that state variables cannot change instantaneously.
Therefore, the values of the state variable remain zero, at the
instant immediately after the Heaviside step input is applied,
at time t = 0+ . The initial value of the output variable, t = 0+,
is determined by using the values of the input variable and
the energy storage (state) variable at time t = 0+ in the energetic equations.
If the system is energized and running in steady-state at
time t = 0 , under a step input applied at an unknown negative time, ta, the response of the system to additional step inputs can be solved by superposition, by including the steadystate response (the particular solution) of the previous input
to the response function. In general, knowledge of the values
of the input and output variables is not sufficient to calculate
all remaining power variables in a system. However, in the
specific case in which a system was subjected to a prior step
input and has reached steady-state, all power variables in the
system will have reached constant values, since the input is
constant in steady-state. Consequently, all derivatives with
respect to time equal zero. This additional knowledge allows
us to determine the value of the non-state variable of each
energy storage element, giving us enough information to establish the values of the state variables at time t = 0 .
The remaining possibility is that the system is in a transient state at time t = 0 , due to a prior step input. This case
can only be solved by using superposition, if the time at
which the prior step input was applied is known. If it is unknown, then the response must be determined with the initial
value method.
Again, superposition only works with linear systems. If
a system can be reasonably approximated (modeled) using
linear elemental equations, then each input to the system creates an output of the same form, but its amplitude is scaled
by the amplitude of the input step function. When a response
function is formulated as the sum (or superposition) of the

183

systems responses to all inputs received, we must assume


the system to be de-energized for each input, or we will double-count the effect of an input.

Problems
Reminders
1. Write energetic equations with proper notation. The
problem statements do not explicitly state that the problems require (a) the energetic equations and (b) proper
notation, because those are part of the linear graph
method and implied by drawing a linear graph. The
energetic equations consist of compatibility, continuity,
element, and energy equations. Proper notation refers to
using (a) node subscripts to identify the positive direction of the drop in the across variable of an element, and
(b) the element parameter as the subscript, which identifies the through variable of an element.
2. Clear fractions and create common denominators as
you work the reduction. Improper fractions must be
cleared to present the result in standard form. It is generally easier to place a sum of ratios over a common denominator when it is created, rather than carrying improper ratios forward and clearing them at the end of the reduction.
One advantage of placing sums of ratios over a common
denominator is that the resulting ratio can be cleared from
a product by multiplying its fractional inverse.
3. Standard form. (a) Time constant form applies only to
the first-order system equations. The terms are ordered,
with the derivative first, followed by the zero-order
term. (b) Standard form for higher order differential
equations requires the coefficient of the highest-order
derivative of the output variable to be cleared. This is
the same standard form used for polynomials. (c) The
system equation cannot have improper ratios. Improper
ratios are not standard form, because they promote error.
4. Unit checks. Check units of the system equation before
solving it, by expressing element parameters in terms
of the systems power variables and time. Do not check
units in terms of fundamental units. Although fundamental units are straightforward in a system of single energy
type, they become very cumbersome in hybrid systems of more than one energy type, for example, electric
motors. For example, how would you combine mechanical units of torque with electrical units of voltage?
5. Convert units to SI for calculations. Convert result to
US customary units, if required. Remember that metric units may not be SI, for example, centimeter. The SI
unit of length is meter. Likewise, remove scaling prefixes and express values in the base unit, for example,
kN = 1, 000 N .
6. Causality. If the system is de-energized for time, t < 0,
then the response of the system for time t < 0 is zero.

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

184
Fig. P3.1 Input pulses

10

20

F(t), N

v(t)
m 0 0
___
sec

10
0

t, sec

t, sec

-5

-10

20
15

200
100

F(t)
N 0

-100

However, evaluating response functions for negative


time will yield a non-zero result, if the response function is not multiplied by a Heaviside step function time
shifted to when the corresponding input is applied to
the system. Do not plot negative time unless you use a
Heaviside unit step function to zero-out each response,
until its corresponding input is applied to the system,
and the response has the correct time shift.
7. Inspect the plots of your results. Look at your results,
and check that they are reasonable and complete. Is the
response a reasonable shape? Does the plot show a response prior to the application of the input? Is the vertical extent of the trace shown, or must the axis limits be
edited? Are the limits of the vertical axis too far from
the extent of the response, thereby squashing the trace?
Does the plot show the beginning of steady-state but not
excessively so?
8. Title plots and label axes. Unidentified plots and unlabeled axes are unacceptable.
9. Mathcad plots. When an xy plot is inserted in a
Mathcad worksheet, Mathcad automatically sets the
ranges of both axes. Mathcad always sets the limits of
the abscissa (the independent variable) from 10 to +10.
Click on a limit to edit it. Set the lower limit of time to
zero, and set the upper limit, such that the response has
reached steady-state. Show the entire transient period of
the response but not too much of steady-state. Edit the
limits of the ordinate to maximize the proportion of the
plot occupied by the trace. Right click on the plot to
bring up the Format menu. Add gridlines to the two

10

15

t, sec

10
F(t)
5
kN
0
-5

10

20

t, sec

30

-10

axes. Select Traces in the Format menu to change the


thickness, color, and type of line.
10. MATLAB plots. Show the entire transient period of
the response but not too much of steady-state. Edit the
limits of the ordinate to maximize the proportion of the
plot occupied by the trace. Add gridlines and change the
trace thickness.
Problem 3.1 For the pulses shown in Fig.P3.1:
3.1.a Sketch the time shifted and scaled Heaviside unit step
functions which superpose to form the pulse.
3.1.b Express the pulse as a function consisting of time
shifted and scaled Heaviside unit step functions.
3.1.c Using Mathcad or MATLAB.
i Plot the time shifted and scaled Heaviside step
functions sketched in part a on the same plot.
ii Plot the pulse function derived in part b.
Problem 3.2 A translational mechanical system consisting of
a force source, a mass, M = 5 kg supported on ideal frictionless rollers, and damper, b = 3 N sec/m , is shown schematically in Fig.P3.2a and b.
3.2.a Use the information presented in the schematics and
linear graph shown in Fig.P3.2 to derive a complete
set of energetic equations for the system.
3.2.b Derive the system equation that relates the force input
to these variables:
i The velocity of the mass.
ii The force acting to accelerate the mass.
iii The force acting through the damper.

Problems
Fig. P3.2a System with a
translational force source, mass,
and damper system. b Schematic
annotated with nodes of distinct
values of velocity. c Linear graph
of the system. d Force input
pulse which acts on the system

185

x,v

F(t)

F(t)

x,v

20

F(t)

F(t), N

10
0

g
Fig. P3.3a System with a translational spring, damper system
acted on by a velocity source.
b Schematic annotated with
nodes of distinct velocity.
c Linear graph of system.
d Velocity input pulse which
acts on the system

x,v

v(t)

x,v

d
2

10
5

v(t)

v(t)
g

t, sec

v(t)
m 0 0
___
sec

t, sec

-5

g
Check the units of the system equation in terms of power
variables and time.
3.2.c The system is at rest at time, t<0. Determine the unit
step responses of the system equations derived in Part
b. Use superposition to determine the responses to the
input force pulse, F(t), plotted in Fig.P3.2d.
3.2.d Plot the responses, using Mathcad or MATLAB.
Problem 3.3 A translational mechanical system consisting of a velocity source, a spring with spring constant,
K = 4 N / m , and a dashpot (damper) with damping coefficient, b = 2 N sec / m , is shown in the schematics, as given
in Fig.P3.3a and b.

-10

3.3.a Use the information presented in the schematics and


linear graph shown in Fig.P3.3 to derive a complete
set of energetic equations for the system.
3.3.b Derive the system equation that relates the applied
velocity input to
i The force in the spring.
ii The velocity drop across the spring.
Check the units of the system equations, in terms of power
variables and time.
3.3.c The system is relaxed at time, t<0. Determine the
unit step responses of the system equations derived in
part b. Use superposition to determine the responses to
the applied velocity pulse, v(t), plotted in Fig.P3.3d.
3.3.d Plot the responses, using Mathcad or MATLAB.

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

186
Fig. P3.4a System with a
translational spring, damper
system acted on by a velocity
source. b Schematic annotated
with nodes of distinct velocity.
cLinear graph of system.
dVelocity input applied to
the system

x,v

v(t)

d
2

v(t)
g

x,v

cm
v(t), ___
sec
10 = v0-

v(t)

t, sec

-10 = v0+

Problem 3.4 A translational mechanical system consisting of a velocity source, a spring with spring constant,
K = 2 N/m, and a dashpot (damper) with damping coefficient, b = 4 N sec/m, is shown in the schematics, as given in
Fig.P3.3a and b. The velocity, 10 cm/sec, was applied at an
unknown time t < 0. The system reached steady-state by time
t = 0 . At time t = 0 , the velocity applied to the system is reversed, v(t)=10cm/sec, for t>0, as shown in Fig.P3.3d.
3.4.a Use the information, as presented in the schematics
and linear graph shown in Fig.P3.4, to derive a complete set of energetic equations for the system.
3.4.b Derive the system equation that relates the applied
velocity input to
i The force in the spring.
ii The velocity drop across the spring.
Check the units of the system equations, in terms of power
variables and time.
3.4.c Use the initial condition method to determine the
response of the systems power variables from Part b to
the velocity applied at time t = 0, shown in Fig.P3.4d.
3.4.d Plot the responses using Mathcad or MATLAB.
Problem 3.5 A translational mechanical system which consists of a velocity source, a mass, and a damper is shown in
Fig.P3.5. Its mass is M = 5 kg, supported on ideal, frictionless
rollers. The damping constant is b = 7 N sec / m. The system
was at rest before the input shown in Fig.P3.5d was applied.
3.5.a Derive the system equation that relates the applied
velocity input to
i The force acting to accelerate the mass.
ii The velocity of the mass.
iii The velocity drop across the damper.
Check the units of the system equations in terms of power
variables and time.

3.5.b The input velocity, v(t), plotted in Fig.P3.5d is


applied to the damper at time t = 0. Determine the unit
step response of the system equations, derived in part
a. Use superposition to solve the system equations for
the velocity input, as shown in Fig.P3.5d.
3.5.c Plot the responses using Mathcad or MATLAB.
Problem 3.6 A translational mechanical system consisting of
a force source, a spring with K = 100 N / m, a damper with
b = 300 N sec / m, and a massless, rigid bar constrained to
translation is shown in Fig.P3.6a. The system has reached
steady-state under the application of an input of 40 N,
applied at an unknown time, t < 0. At time t = 0, the input
force is increased to +60N, as shown in Fig.P3.6d.
3.6.a Derive the system equation which relates the input
force F(t) to
i The force acting through the spring, FK.
ii The force acting through the damper, Fb.
iii The velocity of the point of application of the
input force, F(t).
Check the units of the system equations in terms of power
variables and time.
3.6.b Use the method of undetermined coefficients to solve
the system equations of part a, for the input shown in
Fig.P3.6d.
3.6.c Plot the response using Mathcad or MATLAB.
Problem 3.7 A translational mechanical system which consists of a force source, a mass, a spring, and a damper is
shown in Fig.P3.7. The system was de-energized, before
the input shown in Fig.P3.7d was applied. The parameter values are Case I: b = 800 N sec/m , M = 250 kg, and
K = 400 N/m and Case II: b = 400 N sec/m, M = 250 kg,
and K = 800 N/m .

Problems
Fig. P3.5a System with a
velocity source, damper, and
mass. b Schematic annotated
with nodes of distinct values of
velocity. c Linear graph of the
system. d Velocity input applied
to the system

187

v(t)

g v(t)

M
g

v(t)

10

v(t)
0
0
m
___
sec

t, sec

-10

g
Fig. P3.6a Translational
mechanical system. b Schematic
annotated with nodes of distinct
values of velocity. c Linear
graph of the system. d Force
input applied to the system

x,v

x,v

K
F(t)

K
F(t)

F(t), N
F(0+)

F(t)

= 60 N

K
0

t, sec

F(0 -) = -40 N

g
3.7.a Derive the system equation that relates the applied
velocity input to
i The force acting to accelerate the mass.
ii The velocity of the mass.
iii The force acting through the spring.
Check the units of the system equations in terms of power
variables and time.
3.7.b The input force, F(t), plotted in Fig.P3.7d is applied
to the damper at time t = 0 . Determine the unit step

response of the system equations derived in part a.


Use superposition to solve the system equations for
the velocity input, shown in Fig.P3.7d.
3.7.c Plot the responses using Mathcad or MATLAB.
Problem 3.8 A translational mechanical system which consists of a force source, a mass, a spring, and a damper is
shown in Fig.P3.8. The system was in steady-state under
a previous step input, before the force pulse shown in

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

188
Fig. P3.7a System with a
force source, mass, spring, and
damper. b Schematic annotated
with nodes of distinct values of
velocity. c Linear graph of the
system. d Force input applied to
the system

x,v

b
F(t)

x,v

F(t)

K
Frictionless rollers

Frictionless rollers

F(t)

F(t)
kN 2
0

x,v

b
F(t)

10

t, sec

x,v

F(t)

K
Frictionless rollers

Frictionless rollers

15

-2

g
Fig. P3.8a System with a
force source, mass, spring, and
damper. b Schematic annotated
with nodes of distinct values of
velocity. c Linear graph of the
system. d Force input applied to
the system. The system was in
steady-state under a previous step
input at time, t = 0, when pulse
input, F(t), was applied

F(t)

F(t)
kN 2
0

0
-2

10

15

t, sec

Fig. P3.8d was applied. The parameter values are Case I:


b = 400 N sec/m , M = 250 kg, and K = 5, 000 N/m and
Case II: b = 400 N sec/m, M = 25 kg, and K = 1, 200 N/m.
3.8.a Derive the system equation that relates the applied
velocity input to
i The force acting to accelerate the mass.
ii The velocity of the mass.
iii The force acting through the spring.

Check the units of the system equations in terms of power


variables and time.
3.8.b The input force, F(t), plotted in Fig.P3.8d is applied
to the damper at time t = 0. Determine the unit step
response of the system equations, derived in part a.
Use superposition to solve the system equations for
the velocity input, shown in Fig.P3.8d.
3.8.c Plot the responses using Mathcad or MATLAB.

Problems
Fig. P3.9a System with a velocity source, damper, mass, and
spring. b Schematic annotated
with nodes of distinct values of
velocity. c Linear graph of the
system. d Velocity input applied
to the system. The system was
in steady-state under a previous
step input at time, t = 0, when the
pulse input, v(t), was applied

189

x,v

v(t)

x,v

g v(t)

100
50

v(t)
cm 0 0
___
sec

v(t)

10

15

t, sec

-50

g
Fig. P3.10 a System with
a velocity source, damper,
mass, and spring. b Schematic
annotated with nodes of distinct
values of velocity. c Linear
graph of the system. dVelocity
input applied to the system. The
system was in steady-state under
a previous step input at time,
t = 0, when pulse input, v(t),
was applied

-100

x,v

v(t)

x,v

g v(t)

100

v(t)
cm 0 0
___
sec
-50

Problem 3.9 A translational mechanical system which consists of a velocity source, a damper, a mass, and a spring is
shown in Fig.P3.9. The system was in steady-state under
a previous step input, before the velocity pulse shown in
Fig. P3.9d was applied. The parameter values are CaseI:
b = 600 N sec/m , M = 150 kg, and K = 500 N/m and
CaseII: 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
velocity input to
i The force acting to accelerate the mass.
ii The velocity of the mass.
iii The force acting through the spring.
Check the units of the system equations in terms of power
variables and time.

50

v(t)

Frictionless rollers

Frictionless rollers

Frictionless rollers

Frictionless rollers

10

15

t, sec

-100

3.9.b The input velocity, v(t), plotted in Fig.P3.9d is applied


to the damper at time t = 0. Determine the unit step
response of the system equations derived in part a.
Use superposition to solve the system equations for
the velocity input, shown in Fig.P3.9d.
3.9.c Plot the responses using Mathcad or MATLAB.
Problem 3.10 A translational mechanical system which consists of a velocity source, a damper, a mass, and a spring is
shown in Fig.P3.10. The system was in steady-state under
a previous step input, before the velocity pulse shown in
Fig.P3.10d was applied. The parameter values are CaseI:
b = 400 N sec/m , M = 250 kg, and K = 5, 000 N/m and
CaseII: b = 800 N sec/m, M = 50 kg, and K = 2, 000 N/m.

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

190

3.10.a Derive the system equation that relates the applied


velocity input to
i The force acting to accelerate the mass.
ii The velocity of the mass.
iii The force acting through the spring.
Check units of the system equations, in terms of power variables and time.
3.10.b The input velocity, v(t), plotted in Fig.P3.10d is
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
input shown in Fig.P3.10d.
3.10.c Plot the responses using Mathcad or MATLAB.

Chapter 3 Appendix
Mathcad: Plotting Superposed Functions
Superposed, scaled, and time shifted step responses can be
plotted in Mathcad, either by cutting, pasting, and editing the
response function or by defining a function of two variables:
time, t, and time shift, ts. We will use the input, F(t), and
response function, Fb(t), of Sect.3.8.1 as examples,
(3.61)
F (t ) = 10 N us (t ) 20 N us (t 1) + 10 N us (t 2)


Fb (t ) = 1 N e

K
t
b

10 us (t ) 1 N e

+1N e

K
(t 2 )
b

K
(t 1)
b

20 us (t 1)

10 us (t 2)

(3.62)

Evaluating Eq.3.62 for the parameter values, K = 5, 000 N/m


and b = 2,500 N sec/m , yields

+ 1 N e 2(t 2)10 us (t 2)
Editing, cutting, and pasting a function in Mathcad require
attention to the location and extent of the L-shaped insertion
cursor. The input function assignment statement is the sum
of three Heaviside step function, and Mathcads notation for
the Heaviside step is ()
F ( t ) := 10 ( t ) 20 ( t 1) + 10 ( t 2)
Fb ( t ) := e 2t 10 ( t ) e 2 (t 1) 20 ( t 1)
+ e 2 (t 2) 10 ( t 2)
The subscript b of Fb ( t ) is a literal subscript, created by typing a period after F, as appears here, Fb(t).

Fb2 ( t,ts ) := e 2 (t ts) ( t ts )


The response function is the sum or superposition of Fb2 ( t,ts)
with the same scaling factors used for the input,
Fb ( t ) := Fb2 ( t,0)10 Fb2 ( t,1) 20 + Fb2 ( t,2)10

MATLAB: Plotting Superposed Functions


We will revise the script, Plotting.m, from the Appendix of
Chap.2 to plot superposed, scaled, and time shifted step response functions. The duration of the plot is now the sum of
the duration of the input function plus a minimum of five
times the maximum time constant of the system to allow the
response to reach steady-state after the pulse input ends. If
there is an existing step input acting on the system prior to
time t = 0 , then we must add an additional duration of five
time constants.
if statement
There is no Heaviside step function in MATLAB, because
there is no need for one. We will implement the equivalent
of a Heaviside step function, by using an if statement. If
statements permit the conditional execution of a block of
code. The syntax of an if statement varies with the programming language, but it is a fundamental instruction in all programming languages. The essence of an if statement is as
follows:
if (expression A Relational Operator expression B) is
true then execute the instructions which follow.
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)

The alternative method to cutting and pasting is to define


a function of two variables, time, t, and the time shift, ts,

Instruction executed if conditional is true


end
The instructions between the conditional statement with the
relational operator and the end statement are executed if the
conditional statement is true. If the conditional statement is
false then execution (or control) skips to the instruction
following the end statement. Important: MATLAB is casesensitive. The keywords, if and end, are all lowercase.
MATLAB highlights the leading if and the closing end in
blue to demarcate the limits of the instructions which will
execute, if the conditional statement is true. To further demarcate the if statement as a block of code, it is good practice to indent the instructions which are executed by the if
statement.

Superposition Using Nested Loops

191

MATLAB Time Shift

MATLABs relational operators are listed in Table8.6,


reproduced below.

<

Less than

>

Greater than

<=

Less than or equal to

>=

Greater than or equal to

==

Equal

~=

Not equal

Note the use of a double equal sign for the relational operator, Equal. MATLAB uses a single equal sign as its assignment operator. Also note the symbol for Not Equal, ~=. Most
programming languages use an exclamation point as the
Not or logical inversion operator. However, MATLABs
symbol, ~=, reads as approximately equal, which is why
MATLAB chose it to represent Not Equal. The relational
operator Equal fails unless the two variables or expressions
compared are not exactly equal. For this reason, the Equal
operator should not be used with floating point (scientific
notation) variables or expressions. Floating point variables
carry so many figures it is unlikely that two variables will be
exactly equal after any significant amount of computation.
The Equal operator should only be used with MATLABs
logical (Boolean) variables or with integer variables or expressions. MATLABs logical variables can have one of two
values: true or false and equivalently 1 or 0. MATLAB is
case-sensitive; true and false are all lowercase.

Time shift
The MATLAB equivalent of a Heaviside step function is the
following if statement, where t is the time vector, ts is the
time shift in seconds, and Us is the variable, which plays the
role of the unit step. Us is initialized with the value of zero
before its use.

Us = 0
if (t(n) - ts >= 0);
Us = 1;
end;
This logic is illustrated in the following code, which plots a
stable exponential growth with a time shift of 3 sec.
% TshiftPlot.m
% Time shifted step response plotting script
% Initialize unit step variable Us to zero
Us = 0
% Time Constant
tau = 0.5

Amplitude

Table 8.6 MATLABs relational operators

0.8
0.6
0.4
0.2
0
0

time, seconds

Fig. A3.1 Time shifted response function plot created with TshiftPlot.m
and then formatted

% Time shift ts. Positive value for shift into t>0.


ts = 3
% Duration of calculation equals time shift + seven
% time constants
tmax = ts + 7 * tau
% Time Step
dt = tau/200
% Number of iterations
N = tmax/dt
% Beginning of for loop
for n=1:N
t(n)=(n-1)*dt;
% Beginning of if statement
if (t(n) - ts >= 0);
Us = 1;
end;
% End of if statement
% Response function
y(n) = Us * (1 - exp(-(t(n)-ts)/tau));
end
% End of for loop
plot(t,y)

Superposition Using Nested Loops


The superposition can be performed, using nested loops or
repeated function calls. We will defer the topic of userdefined functions in MATLAB until Chap.8. Briefly, a function is a script with an argument list, allowing it to be used
in MATLAB instructions in the same manner as familiar
mathematical functions, such as sin().
Nested loops are two or more loops, where an inner
loop completely iterates through the range of its counter
variable for each single iteration of the outer loops counter
variable. Fig.A3.2 is a flowchart of a nest loop which initializes a three by four array with ones.

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

192

Enter Outer Loop


of the Nested Loops

Start

Assign values to scalar variables


tau_max, tchar, and nsteps

r=0

Assign values to vector variables


ts( ) and In( )
for loop to initialize vector variable
Us(m)=0, for m= 1 to nsteps

r=r+1
c=0
Enter Inner Loop

Calculate duration tmax, time step dt,


and number of interations N
Initialize outer loop counter
n=0

c=c+1
Array(r,c) = 1

No

Enter Outer Loop

n = n+ 1
t(n) = (n-1)*dt
y(n) = 0
m=0

c == 4 ?

Enter Inner Loop

Yes
Exit Inner Loop
No

m=m+1

No

r == 3 ?

t(n) - ts(m) >= 0 ?


Yes

Yes
Exit Outer Loop

Us(m) = 1

Fig. A3.2 Flowchart of nested loops which creates a three by four


array named Array filled with ones

We will use a similar structure of two nested loops to perform superposition of scaled and time shifted response functions, to plot the response function for a pulse input. We will
demonstrate the procedure using the response of the damper
force, Fb(t), of the force source spring-damper system of
Fig.3.54 to the input function, Eq.3.59. The response function, Fb(t), Eq.3.62, evaluated using K = 5, 000 N/m and
b = 2,500 N sec/m is the following:

y(n) = y(n) + Us(m)*In(m)*StepResponse(t(n) - ts(m))

No

m == nsteps ?
Yes

Exit Inner Loop

No

n == N ?
Yes

Exit Outer Loop

plot(t,y)

 Fb (t ) = 1 N e

K
t
b

10 us (t ) 1 N e

+1N e

K
(t 2 )
b

K
(t 1)
b

10 us (t 2)

20 us (t 1)

(3.62)

The pulse response plotting program is structured as shown


in the flowchart, Fig.A3.4. The program begins with assignment statements which define the maximum time constant, the minimum characteristic times of the system, and
the number of step inputs, nsteps. In the example of a firstorder step response, there is only one characteristic time,

Fig. A3.3 Flowchart of MATLAB script, SuperposedResponse.m

the time constant, and it is, therefore, the value of both


variables tau_max and tchar. Next, the vector variables for
time shifting and scaling the input functions Heaviside step
function are assigned values. The vector of time shifts is
named ts. The vector of scaling factors is named In. A vector variable named Us will play the role of the Heaviside
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

10

Fb(t), N

5
0
5

10
15
20

0.5

1.5

t, sec

2.5

3.5

Fig. A3.4 Responsefunction Eq.3.62 plottedwith MATLAB script,


SuperposedResponse.m, and then formatted

duration of the calculation, tmax, depends on both the maximum time constant of the system and the duration of the
input function.
The outer loop increments through the time step of the
calculation, as in the case of a single step response. It begins
by calculating the current value of the time variable, t(n),
and initializing the current value of the output variable, y(n),
to zero. Control then passes to the inner loop, which increments through the number of step inputs in the pulse input
function. The inner loops if statement checks the time shift,
ts(m), against the time variable, t(n). If the unit step should
transition from zero to one, the unit step variable, Us(m), is
assigned the value of one. If the test fails, Us(m) retains its
initialization value of zero. The product of the input scaling
factor, In(m), the unit step variable, Us(m), and the unit step
response function evaluated for t(n) with the time shift ts(m)
is summed to the output variable, y(n).
The MATLAB script, SuperposedResponse.mfollows.
% SuperposedResponse.m
% Superposition Plotting Script
%
% Maximim Time Constant
tau_max = 0.5
% Minimum Characteristic Time
tchar = 0.5
% Number of steps
nsteps = 3

% Input time shift ts(m). Positive value for shift into t>0.
ts(1) = 0
ts(2) = 1
ts(3) = 2
% Input scaling factors
In(1) = 10
In(2) = -20
In(3) = 10
% Initialize unit step vector Us(m)to zero
for m=1:nsteps
Us(m)=0
end
% Duration of calculation equals time shift +
% seven time constants
tmax = ts(3) + 7 * tau
% Time step
dt = tchar/200
% Number of iterations of the for outer loop
N = tmax/dt
% Beginning of the outer for loop
for n=1:N
t(n)=(n-1)*dt;
y(n)=0;
% Beginning of the inner for loop
for m=1:nsteps
% if statement to transition unit step variable Us(m)
if (t(n) - ts(m) >= 0);
Us(m) = 1;
y(n) = y(n) + Us(m) * In(m) * ( exp(-(t(n)-ts(m))/0.5 ));
end;
% End of if statement
end;
% End of inner for loop
end
% End of outer for loop
plot(t,y)

References and Suggested Reading


Hildebrand FB (1976) Advanced calculus for applications, 2nd edn.
Prentice-Hall, Englewood Cliffs
Rowell D, Wormley DN (1997) System dynamics: An introduction.
Prentice-Hall, Upper Saddle River
Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
dynamics. Addison-Wesley, Reading

Mechanical Systems

Abstract

A mechanical system is a system in which the dominant forms of energy storage, transfer,
and dissipation are described by Newtons laws. Fluid systems are described by Newtons
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 polymers and biological tissues, are described by the dynamic response.

4.1Translational Mechanical System


Elements
Translational mechanical system elements were introduced
in Chaps.1 and 2 and then used in Chap.3 to introduce the
linear graph method. This chapter investigates how to estimate the linear element equations from analytical models of
machine components and dynamic tests. Rotational systems
are then introduced and the similarities and differences between translational and rotational elements are investigated.
We begin with a brief review and summary of translational mechanical power, energy, elements, and models.
An energetic model of a mechanical system does not have
a physical resemblance to the system. An energetic model
represents the energetic behavior of the mechanical system
as the power flow in a network of interconnecting elements
which supply, store, or dissipate energy. Only the dominant
aspects of a systems energetic nature are modeled, in part,
not only for economy of effort but also due to the limitation
of our knowledge of a systems actual behavior.
Mechanical systems store energy as kinetic energy (energy of motion) or strain energy (energy of elastic deforma-

tion). Gravitational potential energy will not be represented


as an energy storage mode in translational mechanical systems. Gravity will be represented as a force source acting
on a mass element. Kinetic energy is dissipated as heat
through friction, which is lost from the mechanical system.
We will categorize mechanical systems based on the type
of motion, either translation (linear) or rotation. Further,
we will restrict motion in a subsystem to one dimension, so
that we can work with scalar velocities rather than vector
velocities. It may be that we will need a number of subsystems, each representing motion in one dimension, to represent the motion of a machine.
In translational mechanical systems, power, the flow rate
of energy, is the dot product of force and velocity,


dW
= P = F v
dt

(1.17)

The elemental and energy equations for translational mechanical systems written in terms of the power variables
force and translational velocity are summarized in Table4.1.

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_4, Springer Science+Business Media New York 2014

195

4 Mechanical Systems

196
Table 4.1 Translational mechanical elements
Elemental equation
Energy equation
dv1g

Mass

Fm = m

Spring

dFK
= Kv12
dt

Damper

Fb = bv12

dt

Em =

1 2
mv1g
2

EK =

FK2
2K

None.
Energy dissipater

4.2 Modeling Translational Elements


All engineering calculations are estimates to some degree.
Reasonably accurate estimates of the element parameters
for the elemental and energy storage equations are needed
in order to use dynamic system theory. The difficulty level
in obtaining a reasonably accurate estimate varies with the
type of physical component and attribute to be modeled. Difficulty also varies with the point in the engineering cycle,
i.e., does the physical component exist or is it still in design?
There is also a chicken-and-egg aspect to the computation.
Our lumped models include elements to represent significant energetic attributes of the system, where significance is
defined as the amount of energy stored or dissipated in an attribute relative to the other attributes of the system. We must
first compute the elemental parameter before we can judge
the significance of the element in the system.
When possible, the best estimates of elemental parameters are calculated from the dynamic response of systems.
Characterization of the transient period of first-order step responses yields the time constant. The steady-state value of a
step response yields a coefficient. Likewise, as we will see
when we work with second and higher-order systems, the
transient period of an oscillatory underdamped second-order step responses yields the damped natural frequency and
damping ratio. The steady-state value of both underdamped
and overdamped second-order step responses yield useful
coefficients.
If the component, machine, or system does not yet exist,
parameter values can be estimated using the geometric configuration of the machine or system component, values for
the various material properties, and estimates of the operating range of the power variables. If possible, estimates produced by interpolating or extrapolating parameters obtained
from similar, existing, well-characterized components will
have the least uncertainty. If no data are available from similar components, then estimates must be based on published
material properties and engineering models. The published
values of bulk properties, such as the density of most engineering materials or the elastic properties of steel, can be
used with great confidence. In other cases, material properties can be very uncertain with published values representing

only a limiting case. For example, aluminum does not have a


linear elastic response. The modulus is defined by a line offset from and parallel to the approximate slope of the initial
region of the stressstrain curve. The greatest uncertainties
in material properties tend to be in those governed by flaws
within the material, such as strengths, and the condition of
the surface of the material, such as friction.
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 properties is the yield strength of metal alloys. Yield strengths
for different grades of steel and aluminum are published
in various handbooks, but what do the values represent?
The published yield strength is surely not the average yield
strength. If it were, then half of all specimens would yield
below that value. The mean yield strength is surely greater
than the published yield strength, but that value and the
standard deviation are not available since they would reveal the steel mills control over the manufacturing process, which is proprietary information. Further, strength
is affected by the processing history of the part. Although
the history of a part is well documented in some instances,
such as in aerospace applications, in others, it may be unknown.
We will begin with the simplest approach, calculating
elemental parameters using published material properties,
and then consider how to estimate parameters from experimental data. Most of the materials and devices we work
with are non-linear. We investigate the process of linearization in Sect.4.2.4.1. We shall see that linearizing about
an operating point is only accurate over a limited range
of the systems power variables. If a single straight-line
approximation of a constitutive relationship is not accurate
enough to yield useful results from linear calculations (i.e.,
superposition), then, we must resort to numerical methods so that we can use non-linear models of springs and
friction.
We will defer using non-linear element equations until
we have introduced the state-space representation of dynamic systems in Chap.7 and finite difference methods
in Chap.8. The RungeKutta algorithm, which we shall
program in MATLAB, will solve state-space formulations
which include non-linearities. An alternative approach for
dealing with non-linear energetic elements is the use of dynamic simulation software, such as MathWorks Simulink.
Simulink has a graphical user interface in which the engineer
constructs a block diagram of the dynamic system. Block
diagrams are discussed in Chap.9. Simulink, however, is not
addressed in this text for two reasons. First, instruction in
the use of graphically based engineering software which includes tutorials and examples is unnecessary for an educated
and enterprising engineer. Second, those readers for whom
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


MATLAB programming exercises can be performed using
free, open-source programming language R, which was created to resemble MATLAB. Alternatively, those with some
programming experience may choose to work in a general
purpose open-source programming such as C. Simulink is
not free, in fact, it is very expensive for those who do not
have student status.

4.2.1 Mass, Kinetic Energy Storage Element


An element in a model represents a single energetic property.
All physical components in machines possess mass, but the
property of mass is not assigned to all physical components.
Mass elements are added to the model of a dynamic system, if the mass of a component (or assembly) accounts for
a significant amount of power flow or energy storage within
the system, relative to power flows and energy storage in
other modes. A significant power flow can be created by a
large acceleration of a modest mass. Likewise, a significant
amount of kinetic energy can be stored in a small mass at
high velocity. Consequently, the nature of the system and the
magnitude and rate of change of its power variables, in other
words, the systems operating conditions, must be considered when modeling a system.
F = ma is the familiar form of Newtons second law.
More formally, Newtons second law equates force and the
rate of change of momentum,
F=

dp
dt

F=

dmv
dt

If mass m is constant, then


F=m

dv
dt

Restricting motion of a single dimension (or direction) allows this vector equation to be written in scalar form
F=m

dv
dt

As you know from your study of physics, as part of his theory of relativity, Einstein revised Newtons second law to
correct for the velocity dependence of mass
m=

m0
1

v2
c2

where mass m0 is the rest mass and c is the speed of light


in a vacuum, 300,000km/sec. Subatomic particles can move

at speeds which are substantial fractions of the speed of light.


Macroscopic matter does not. The maximum velocity an engineer may deal with is the escape velocity from earth orbit
of approximately 11km/sec. The relativistic increase in mass
at that escape velocity is vanishingly small
m0

m=
1

km
11
sec

km

300, 000

sec

m0
1 1.34 10 9

m = m0

Although the true and non-linear elemental equation for


mass is relativistic, Newtons second law is extremely accurate for all practical purposes above the atomic scale.
Although the element equation of a mass is linear, it may
still be uncertain due to uncertainty in the parameter value,
i.e., the amount of mass. In practice, we are often unable to
weigh a machine part to determine its mass because we are
not allowed to disassemble the machine or the part is too
large to weigh. We must calculate the objects mass from its
geometry and density. Variation of the part geometry from
that specified in the drawings or, if drawings are not available, the accuracy of our measurements, is a source of error.
Often, the material or the specific alloy is unknown. The uncertainty of the estimated mass depends on how variable the
densities of the alloys of the base material are. The density
of steel can be assumed with certainty. However, the densities of some metal alloys and most polymers vary with the
specific formulation. Finally, when the system is open, in the
sense of allowing mass flow across the system boundary, the
accuracy of the measurement of the mass flow rate limits the
precision with which we know the mass of the system.
The precision of our result is no greater than the least
precise datum used in the calculation. However, precision
is usually the lesser of our concerns. After we take all our
measurements and calculations, a final uncertainty in the parameter value remains in how representative was the part of
the population from which it was drawn. The uncertainty or
variability in the values used in calculating the parameters
has a greater effect on limiting the precision we report for
our result. We must decide how many significant figures of
the result we believe and report no more than those.
Mass is the easiest energetic attribute to calculate when
a machine component is (1) in translational motion, (2) can
be reasonably modeled as rigid (i.e., all points on the component have the same velocity), and (3) the density/ies of the
material(s) which comprise the part are known.
Sadly, a common error is made when computing the mass of
a machine component. Mass is not weight. Mass density does
not equal unit weight. Weight is a force. Specifically, weight is
the force of gravity acting on mass at sea level on earth

4 Mechanical Systems

198

F = Ma M =

Table 4.2 Density variation of aluminum alloys


Aluminium alloys Density (kg/m3)
Lowest
Highest
Percent variation (%)
2000 Series
2,580
2,850
10.5
2.2
6000 Series
2,680
2,740
Casting alloys
2,650
2,950
11.3

F
a

and
weight = Mg
where g, the gravitational constant, is the acceleration of
gravity at sea level
g = 32.2

ft
m
= 9.81
2
sec
sec 2

Pound mass is a unit used to perform fluid mechanics and


thermodynamics calculations in the US customary units. A
pound mass is the mass of a material which weighs a pound
force if the mass is not accelerating. Students who have a
taken the short-sighted formulaic approach to their engineering studies become confused by pound mass because they
do not understand why the gravitational constant is included
in the calculations. The confusion is compounded by the variety of peculiar units used in the US customary units to express
weight and gravitational acceleration. There is no pound mass
in dynamics and system dynamics. Pound is a unit of force.
The unit of mass in the US customary units is a slug
weight = Mg slug

weight
g

where weight is in pounds and the gravitation constant


g = 32.2 ft /sec 2 . The units of a slug are:
weight
[slug ] = g

lb
slug
=
[ ] ft
sec 2

lb sec 2

ft

[slug ] =

It is far easier to avoid wrestling with the US customary units


and convert to SI units at the beginning of the calculation
and then convert back to the US customary units at the end,
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 density is available, it is often the case that only the type of materialbut not the particular alloy metal or type of plasticis
known. For example, say a metallic material is identified as
aluminum. Aluminum, a low density metal, = 2700 kg/m3,
is commonly alloyed with copper, a very dense metal,
= 8960 kg/m3. As a result, aluminum alloys vary significant in density, Table4.2.

What should an engineer do when the aluminum alloy


is unknown? Play the odds. Assume the density of an alloy
commonly used for the application. For example, if one
were calculating the mass of an aluminum component which
was machined, as opposed to cast, then Aluminum 2024,
2024 = 2780 kg/m3, or Aluminum 6061, 6061 = 2700 kg/m3 ,
would be reasonable guesses, yielding a 3% uncertainty. It is
a mistake to report a result with more significant figures than
can be justified by the certainty of the information available.
For the 2000 Series alloys, one certain figure and one estimated figure is reasonable. If you believe the alloy to be
2024, then use its density to three significant figures in your
calculation, but only report your result to two.
Actual machine components are not rigid. When a part is
loaded in compression or tension within its elastic response,
the deformation due to elastic strain is small unless the part
is very large. Although there is kinetic energy stored during elastic axial deformation, it is usually negligible. When
a part is loaded in elastic bending, the maximum displacement and the kinetic energy stored during bending of a small
part can be significant. A tuning fork is an example. Another
example is a dynamic experiment performed by all students
early in their education by bending and then releasing a ruler
cantilevered from a table. The resulting vibration or oscillation requires both strain and kinetic energy storage in the
single physical object, the ruler. The displacement and velocity of the cantilever increase with distance from its support.
The lack of rigidity leads to the distributed mass of the ruler
not storing kinetic energy uniformly but as a function of position along the cantilever. The effective mass is calculated
by weighting (no pun intended) the mass at a point by its
displacement or velocity using a shape function. We will
formulate the displacement of a cantilevered beam loaded at
its end in Sect.4.2.3.4, when we investigate elastic energy
in bending and calculate the effective mass of a cantilevered
beam in Sect.4.2.3.2.

4.2.2 Spring, Strain Energy Storage Element


Springs of many different shapes and sizes are used in machine design. Compression and extension coil springs are the
most familiar. Other common springs are torsion coil springs,
leaf springs, and belleville springs. Torsion coil springs are
used in mouse traps and clothes pins. Leaf springs consist
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


washers, also known as disk springs, are washers or disks
deformed into a profile which approximates a fraction of a
sphere. An endless variety of custom-designed coil springs
can be made on spring machines.
Any matter that deforms elastically stores strain energy. If
the amount of strain energy stored in a machine component
is significant relative to the amount of energy stored elsewhere in the system, then the machine components attribute
of strain energy storage is included in the energetic model
of the system. A component of a given size loaded in bending undergoes substantially more deformation than the same
component loaded axially.

4.2.2.1 Translational Spring Constant


The most familiar expression for a translational spring is
(1.6)
FK = K x

x is the conventional notation for the deformation or


change in length of a spring and K is the stiffness, spring
rate, or spring constant. It is important to keep the distinction between position, displacement, and deformation clear.
A displacement is a change from an initial position. Knowing the displacements of both ends of the springs allows us
to calculate the deformation of the spring.
The reciprocal of stiffness is called compliance
1
compliance
K

where

deformation
force

[compliance] =

(4.1)
Compliance is often given the symbol C. This is unfortunate
because C is also used in place of b as the symbol for proportional or linear damping in vibration analysis.
Translational spring constants can be calculated if the geometry of the machine elements and the moduli of the materials are known or if a reasonable estimate is available. Most
materials do not have a true Youngs modulus, since their
stressstrain relationships are non-linear. Steel is an important
exception, since there is a linear region of the stressstrain
curve below the yield stress. In contrast, the slope of the initial
portion of an aluminum alloys stressstrain curve is used in
the 0.2% offset construction to establish a yield stress.
Although the moduli of metals may be only linear
approximations, the stressstrain behavior of metals below
yield is considered a bulk property and insensitive to microscopic material flaws and common manufacturing operations. Hence, one can place greater confidence in published
values of moduli of metals than in, say, values of ultimate
strength.

Elastic Deformation

F(t)

F(t)

Cross-sectional area A
Youngs modulus E

Fig. 4.1 Axially loaded element with elastic deformation


Table 4.3 Moduli of carbon and alloy steels (excluding stainless steels)
Youngs Modulus

28.5 to 30.0 106 psi

Shear Modulus

11.0 to 11.9 106 psi

Calculation of a translational spring constant is an exercise in deriving a forcedisplacement or forcedeflection


relationship for a machine part of assembly. The forcedisplacement relationship is then expressed as the ratio


Force
Force
or
= Spring Constant K
Displacement
Deflection
(4.2)

4.2.2.2 Axial Strain


Calculation of the spring constant of an element loaded axially in tension is an application of the definition of linear
strain, Fig. 4.1. Compressive loading must always be considered as a critical calculation due to the possibility of a
catastrophic buckling failure, which is outside the scope of
this text.
For tensile axial loads below yield and compressive axial
loads below buckling, the extension or compression of an
axial member is the product of the length of the member and
the linear strain.
Elastic Deformation L = L =
The ratio


L=

Faxial L
AE

Force
yields the spring constant K
Displacement
F
Force
AE
= axial =
K
Displacement
L
L

(4.3)

We often know very little about the materials in a system


we need to analyze or modify and must use published material properties in our calculations of the energetic properties.
Fortunately, the Youngs modulus and shear modulus values
of carbon steel vary relatively little as a function of carbon
content, processing (i.e., cast vs. cold rolled), and heat treatment, Table4.3.

4 Mechanical Systems

200
Fig. 4.2a Linear bending stress
distribution b Infinitesimal force
and moment

-z

Neutral
Axis

+z

4.2.2.3 Deflection in Bending


Forces acting transverse to a member induce a bending moment. The deflection of the point of application of the force
is calculated by beam theory. In contrast to deflection under
axial forces, the deflection due to bending does not depend
of the cross-sectional area of the beam. It depends on the
area moment of inertia I, a purely geometric quantity.
The connections, or support conditions, at the ends of an
element in bending greatly influence the magnitude of the
deflection of the point of application of the force. In order
to calculate bending deflection manually, the support conditions must be statically determinant. This is generally not the
case in machine design, where statically indeterminate connections are stiffer and can carry greater loads because they
have lower local stresses.
Modeling a statically indeterminate machine element in
bending as if it were a statically determinant beam will typically overestimate the bending deflection, yielding a spring
constant which is smaller than the actual. Conversely, the
energy stored in bending strain will be overestimated because of the inverse relationship between the spring constant and strain energy storage. Manual calculations are still
useful as initial estimates and may suffice, if other energetic
parameters of the system are as uncertain. However, when
more refined estimates of the spring constant of a component with indeterminate end conditions or non-uniform
cross-section are warranted, then finite element calculations
are necessary.
A brief review of elastic beam theory follows. The assumption that planar cross-sections of a member remain
planar when the member is bent allows the radius of curvature of the member to be related to the rotation of a plane
section and, consequently, to the axial strain distribution
across the beam in the y-direction. If the material model
is linearly elastic, i.e., = E , the assumption that plane
sections remain plane leads to a linear variation in axial
stress with distance from the neutral axis. Moment equilibrium of a free body cut from the beam results from the

x
dy

dFz = z dA
dM x = ydFz = yz dA

axial stress distribution. The area moment of inertia combines dependence on distance from the neutral axis of both
the linear axial stress distribution and the resisting moment
created by that stress acting on an elemental area of the
cross-section.
The deflection equations can be derived from the relationships for linear bending stress


z = z max
ymax

(4.4)

infinitesimal internal force


dFz = z dA
(4.5)
and infinitesimal internal moment


(4.6)

dM x = ydFz

Expressing the infinitesimal internal moment in terms of the


linear bending stress distribution yields, Fig.4.2


z max
y
dM x = y z max
dA
=

ymax
ymax

2
y dA

(4.7)

where the area moment of inertia is




I = y 2 dA

(4.8)

The assumption of a linear stress distribution allows the moment to be related to the radius of curvature in terms of the
Youngs Modulus, E, and the area moment of inertia of the
cross-section, I


1
rcurvature

M ( x)
EI

(4.9)

4.2 Modeling Translational Elements

FR

201

L
y

MR

FR

Substituting the expression for the internal moment into


Eq.4.10 and evaluating the integral for a cantilevered beam
with a constant cross-section and Youngs modulus

d2y

dx

dy

Fig. 4.3 Model of a cantilevered beam

MR

x
FR

Mz Mz

Fy

L-x

Fy

Deflection of a beam is calculated by approximating the curvature of the beam for small deflections as
rcurvature

d2y
dx 2

y ( 0) =

(4.11)

4.2.2.4 Cantilevered Beam


The moment on an internal cross-section of the cantilevered
beam shown in Fig.4.3 is found by creating a free body of
the right-hand section of the beam with an imaginary cut a
distance (Lx) from the end, Fig.4.4, and then using moment equilibrium at the cut

dy
=0
dx

03
F L02

+ C1 0 0 + C2 = 0

EI 2 0 3 0

dy ( 0)
dx

Integrating Eq.4.11 twice with respect to x yields the displacement y(x) of the beam. Evaluation of the integral requires knowledge of the variation of the bending moment
along the beam M ( x), which is found from statics; hence,
the requirement of statically determinant beams. The two
constants of integration require knowledge of the slope and
displacement somewhere on the beam.
Two of the most commonly used models of translational springs are a cantilevered beam and a simply supported
beam.

M cantilevered ( x ) = F ( L x )

F Lx 2 x 3
+ C1 x + C2

3
EI 2

y ( 0) = 0 and

(4.10)

Equating Eqs.4.9 and 4.10, we get


d 2 y M ( x)
=
EI
dx 2

x2
+ C1 dx
2

The end conditions of a cantilevered beam are that the displacement and slope of the beam are zero at the support

( L x ) dx

x2
Lx

+ C1

y=

Fig. 4.4 Free body diagram of beam cut at distance x from the support

F
EI

dx dx = E I Lx

y
y

dx =

dy
F
=
dx
EI

Area Moment of Inertia I


Youngs Modulus E

FR

F
02

0
L
+ C1 = 0
0

2 0
EI

C2 = 0

C1 = 0

The deflection equation for a cantilevered beam loaded at its


end a distance L from the support is


y ( x) =

F Lx 2 x 3

6
E I 2

(4.12)

The spring constant calculation requires the deflection at the


point of application of the load. Evaluate Eq.4.12 for x = L
y ( L) =


F L3 L3
F 6 L3 2 L3
F L3
=
=

12
3E I
EI 2 6
EI

F
Force
3E I
=
= 3 K Cantilevered Beam
3
Displacement
FL
L
Loaded at End

3E I

(4.13)

4.2.2.5 Simply Supported Beam


Derivation of the deflection equation of a simply supported
beam, Fig.4.5, with a single load is more involved than that
of the cantilevered beam, because the end conditions lead to
the elimination of one of the two constants of integration.
The reaction forces are:

4 Mechanical Systems

202
Fig. 4.5 Model of a simply
supported beam with a single
concentrated load F

F
a

+M

+M

FR
FR

FRA =

F b
F a
and FRB =
L
L

and the moments on cross-sections through the beam are:


M simply

supported

( x ) = FR

x for 0 x a

and
M simply

supported

( x ) = FR ( L x )
B

for a x L

Evaluating the integral for a simply supported beam is


more involved than for a cantilevered beam. The end conditions of a simply supported beam are that the displacement of the beam is zero at the supports. However, the
slope of the beam at the supports is unknown, because the
beam is free to rotate. Hence, only a constant of integration
can be determined by using the displacement of the end
of the beam. It is tempting to assume that the maximum
deflection of the beam occurs at the location of the load,
but that may not be true and cannot be established until
the deflection equation is known. The one exception is the
special case loading the beam at its mid-span, where symmetry requires the slope of the deflected shape to be zero
at the location of the load. In the general case when the
load is not at the mid-span, what we know must be true is
that the deflection of the point of application of the load
and the slope of the beam at that location is the same, regardless of whether we formulate the deflection and slope
equation from support A or support B. Consequently, the
second constant of integration is found by equating the deflection and the slope of the point of application of the
load calculated from both supports. The general case is no
more conceptually challenging than the symmetrical special case but the algebra does become involved.

y
x

Area Moment of Inertia I


Youngs Modulus E

FR

yB

We will evaluate the special case of loading the beam at


mid-span first. Symmetry leads to the two reaction forces
being equal

F RA =

L
2 =F
L
2

and

F RB =

L
2 =F
L
2

L
are derived
2
by integrating the relationship between moment and bending
twice with respect to x
The slope and deflection equations for 0 x

FR
d2y
dx 2 dx = E IA

xdx

for 0 x

L
2

FR x2

dy
= A + C1
dx
EI 2

for 0 x

L
2

F RA

dy

dx dx = E I
yA ( x) =

x2
L
+ C1 dx for 0 x
2
2

F RA x 3

+ C1 x + C2
E I 6

Evaluating the slope equation at x =


L
dy
FR
2
=0= A
dx
EI

for 0 x
L
2

L 2

+ C1
2

L

2
+ C1 = 0
2

C1 =

L2
8

L
2

4.2 Modeling Translational Elements

203

Evaluating the deflection equation at x=0

y A ( 0) = 0 =

y A ( 0) = 0 =

F R A 03
L2
0 + C2

E I 6 0 8

C2 = 0

FR x 3 L2
y A ( x) = A
EI 6 8

for

0 x

Substituting for the reaction force at support A


yMid span ( x) =
Loading

F x 3 L2

2 E I 6 8

0 x

for

L
2

Our objective is the spring constant (or spring rate) of a


beam bent by a transverse load. Consequently, we evaluate
the deflection equation at the point of application of the load,
x = L /2
L 3
F 2
L2
L

yMid span =
2 2E I 6
8
Loading

L


2

F L3 L3
F L3 3 L3
FL3
L
yMid span =
=

2 2 E I 48 16 E I 84 84
48 E I
Loading
Signs are always a problem in mechanics. Negative sign of
the displacement equation is correct relative to the positive
y-direction. The applied force is also in the negative direction, yielding a positive spring constant. Energetic parameters must be positive!

Force
F
48 E I
=
=
K Simply supported
Displacement
L3
F L3
Beam Loaded at

Mid Span
E I 48
(4.14)
The general case of simply supported beam loaded at a
location, x = a, where a L /2 is more involved to derive,
because we cannot assume that the slope of the beam is zero
at the location the load is applied. We begin as we did previously and derive the same slope and deflection equations for
0 x a, where x is measured from support A
FR x2

dy
= A + C1
dx
EI 2

F R x3

y A ( x ) = A + C1 x + C2
EI 6

0 xa

for
for

0 xa

We can evaluate the deflection equation since we know


y A ( 0) = 0

C2 = 0

Hence
yA ( x) =

L
2

F R A 03
+ C1 0 0 + C2

EI 6 0

F RA x 3

+ C1 x

EI 6

0 xa

for

The deflection equation for a x L can be formulated in


terms of x or, more conveniently, in terms of position measured from support B,
d2y

dx =

F RB
EI

dy F RB 2
=
+ C3

d E I 2

FR
dy
d d = E IB
yB ( ) =

2
2

+ C3 d

FR 3

+ C3 + C4
E I 6

0 b

for

0 b

for

for

0 b

for

0 b

Evaluating this deflection equation for = 0, the location of


support B, eliminates one constant of integration
yB ( 0) =
yB ( ) =

F RB 03
+ C3 0 0 + C4

EI 6 0

F RB 3

+ C3

EI 6

for

C4 = 0

0 b

The slope of the beam at this location is independent of how


we chose to formulate the equations. Our formulation written
in terms of approaches the load position from the negative
x-direction. Hence, the slope in terms of is the opposite of
the slope in terms of x. Inverting the sign and equating the
two deflection equations evaluated for the location of the applied load, x = a and = b, yields an equation with constants
of integration, C1 and C3
dy A ( a )
dy (b )
= B
dx
dx

F RA a 2
F RB b 2

C
=

+ C3
+
1

EI 2
EI 2

F b a2
F a b2
+ C1 =
+ C3

EI L 2
E I L 2

a2

b2

b + C1 = a + C3
2

4 Mechanical Systems

204

Evaluating the two deflection equations at the location of the


applied load yields a second equation with C1 and C3
y A ( a ) = yB (b )

yA (a) =

F RA a 3

F RB b3
+
=
+ C3b
C
a
1

EI 6

EI 6

F b a3
F a b3
C
a
+
=
+ C3b
1

EI L 6
EI L 6

yA (a) =

a 2 b ab 2
+
= bC1 aC3
2
2
ba 3 ab3

= abC1 + abC3
6
6

This checks. The result can be expressed in many different


forms because L = a + b
yA (a) =

We can now use Gaussian elimination. Multiplying the first


equation by b and adding it to the second equation eliminates
C3, yielding C1
a 2 b ab 2
+
= bC1 aC3
2
2
a 2 b ab 2
ba 3 ab3

+ b
+
= abC1 + abC3 + b ( bC1 aC3 )
6
6
2
2
a 2 b ab 2
+
= bC1 aC3
2
2
ba 3 ab3 a 2 b 2 ab3

+
+
= abC1 b 2 C1
6
6
2
2

2
2
F a 2b2 F a ( L a )
=

3L
E I 3L E I

yA (a) =

2 2
F ( L b) b

3L
EI

Similarly, the spring constant K can have a number of forms.


Of the above, the latter two are the most convenient for computation


Force
=
Displacement

F
3L E I
= 2 2 K Simply Supported
2 2
ab

Beam Loaded at x = a
F ab

E I 3L
(4.15)

and

a 2 b ab 2
+
= bC1 aC3
2
2
ba 3 ab3 a 2 b 2 1
6 + 3 + 2 ab + b 2 = C1

The deflection equation y A ( x) can now be evaluated for


x=a

for 0 x a

yA (a) =

L L
F
2 2
Fa 2 b 2
F L3
yA (a) =
=
=
3E I L
3E I L
48E I

Form a set of simultaneous equations in terms of C1 and C3

F RA x 3 ba 3 ab3 a 2 b 2 1
+
+
+

3
2 ab + b 2
E I 6 6

Fa 2 b 2
3E I L

We can check this equation against our previous result by


evaluating it for a = b = L /2

a3

b3

b + C1a = a + C3b
6
6

yA ( x) =

2 2
F a 2 b3 + a 3b 2
F a b (b + a )
=

E I L 3 ( a + b ) E I L 3 ( a + b )

( Fb ) a 3 ba 3 ab3 a 2 b 2 1
+
+
+

a
3
2 ab + b 2
E I L 6 6

Force
=
Displacement

F
2
2

F a ( L a)

EI
3L

Force
3L E I
= 2
K Simply Supported
Displacement a ( L a )2
Beam Loaded at x = a

(4.16)

4.2.2.6 Non-Linear Springs


The relationship between force and elastic deformation may
be non-linear, either because the stressstrain properties
of the material are non-linear (e.g., aluminum), or because
the object may deform in a way that alters the stress state
in the object. Consequently, the greatest source of error in
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
5

Nominal
Stress
MPa

x,v

F(t)

Uniaxial

4
3
2

Biaxial

F(t)

Stretch
Fig. 4.6 Stress vs. stretch data for rubber under uniaxial and biaxial
load. Note the unit of stretch rather than strain due to the large deformations. (The data are from Trealor and reproduced from Boyce and
Arruda 2000)

spring constant or spring rate K and not the variation in the


performance of nominally identical springs.
Non-linear springs are often desirable in machine design.
We may wish to have a spring which is soft with a small
spring constant K for small force and becomes stiffer with
a larger K for larger forces. Elastomeric (rubbery) materials
have this property, Fig.4.6. Elastomers are used as vibration isolators to dissipate vibration and as shock mounts
to protect components from impulse loads. An example of
the use of elastomers added to design for their vibration
damping characteristics are motor mounts in automobiles.
Although non-linear springs are useful in mechanical design, they are problematical for the analysis. Superposition
requires linear differential system equations. A linear differential system equation, in turn, requires a linear constitutive
or elemental equation for every element in the system.

4.2.3 Effective Mass


If a machine component stores a significant amount of both
kinetic and elastic energy, then its energetic model must include both a spring and a mass. The spring constant K is
calculated as the ratio of force over displacement, Eq.4.2,
as developed in Sect.4.2.2. The corresponding lumped
parameter mass is the effective mass of the element. A
system with two independent energy storage modes will
have internal energy transfers during its natural or unforced
response. The transfer of energy between independent energy storage modes is manifested as an oscillation in the
systems power variables, which, in a mechanical system,
is termed a vibration. Accordingly, calculation of effective
mass uses theory developed in the mechanical engineering
discipline of vibrations. The study of vibrations and waves
are topics of Newtonian mechanics and, as such, predate the
study of system dynamics by centuries. Lord Rayleigh, a.k.a.

x1g

1
0

b
c

Fig. 4.7 Linear variation in deformation with position along a translational spring with constant stiffness per unit length. a Relaxed spring.
b Compressed spring

x,v

F(t)

v1g

K
b

vag vbg

vcg

Fig. 4.8 Linear variation in velocity with position along spring K during deformation

John William Strutt (18421919), made significant contributions, and common methods bear his name.
An assumption of vibration theory is that the maximum
strain energy stored in an element equals the maximum kinetic energy. In other words, there is no energy dissipation in
the system. The notation used in physics and vibrations is U
for potential energy and T for kinetic energy. The deformed
shape of the object must be established. Three approaches
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. Alternatively, for objects with constant cross-section and properties, such as beams, a sinusoidal shape which satisfies the
end conditions is assumed.

4.2.3.1 Effective Mass of a Spring


When an elastic object undergoes deformation, the displacement of points on the object must vary with position. If all
points on an object displace equally, then the object translates without deformation. In the special case of a machine
component with constant stiffness along its length, the elastic deformation varies linearly from one end to the other, as
shown with spring K in Fig.4.7. The elastic energy stored is
uniform along the length of spring K since it is stored in elastic strain. If the machine component, say spring K, also stores
a significant amount of kinetic energy during the elastic deformation, the velocity of the mass elements which comprise
the component will also vary linearly, Eq.4.17, Fig.4.8.


v( x) = 1

x
v1g
L

(4.17)

4 Mechanical Systems

206
Cross-sectional area A
Area-moment of inertia I
Youngs modulus E
Density

y,v
x

dm

F
1

Fig. 4.9 End-loaded cantilever beam with infinitesimal mass element


dm at x

where L is the instantaneous length of the spring.


Consequently, kinetic energy is not stored uniformly in
spring K. The effective mass is calculated in terms of the
kinetic energy stored in the spring calculated using the maximum velocity, v1g, Eq.4.18:


E kinetic

1
= M effective v12g
2

(4.18)

The effective mass of spring K is the inertia felt at node


1. Since v1g is the maximum velocity of any mass element
in the spring, the effective mass must be less than the total
mass. If the mass of the spring is Mspring, then a mass element
dm of the spring at any instant during deformation is:


dm =

M spring
L

(4.19)

dx

The effective mass is calculated by scaling the contribution


of a mass element dm by the square of the ratio of the velocity at that location to the maximum velocity, Eq.4.19

E kinetic

y ( x) =

Loaded at End

E kinetic =

d 2 y ( x)
dx

1 M spring 2 L
v1g
2 L3
3

E kinetic =

effective

d 3 y ( x)
dx 3

M spring
3

dy ( 0)
=0
dx

M ( x)
d 2 y ( L)

=0
EI
dx 2

Also recall that shear, V(x), integrates to yield moment, M(x),


and, conversely, moment is differentiated to yield shear

1 M spring 2
v1g
2 3

M spring =

(4.13)

The boundary conditions for the free end are the absence of
shear and bending moment. Recall that moment is related to
curvature, which for small deflections can be written as

1 M spring 2 2
x3
v1g L x Lx 2 +
3
2 L
3

3E I
L3

(4.12)

If the external force F is suddenly removed, the cantilevered


beam will vibrate or oscillate, as anyone who has bent and
released a ruler cantilevered off the edge of a table can attest. The vibration occurs due to the transfers in strain energy
from bending into kinetic energy and back. When a cantilevered beam is loaded and then released, its oscillation is
a free, as opposed to a forced, vibration. Free vibrations
are mechanical oscillations described by the homogeneous,
unforced, or natural response of the mechanical system.
The deflected shape of a freely vibrating beam is the superposition of modes of a number (infinite, in the abstract)
of possible deflected shapes, all of which are sinusoidal. The
primary or first mode has the lowest frequency of vibration,
known as the fundamental frequency, and the simplest shape,
with the fewest sinusoidal cycles or fractions of cycles.
The calculation of the effective mass requires assuming
a sinusoidal mode shape. For an assumed mode shape to be
reasonable, it must satisfy the boundary conditions of a cantilevered beam. Where the beam is built-in to a support, the
deflection, y(x), and the slope of the beam, dy(x)/dx equal
zero.
y ( 0) = 0

1 M spring 2
2
v1g ( L x ) dx
3
2 L
0

F0 Lx 2 x 3

6
E I 2

K Cantilevered Beam =

1 M spring 2 2
=
v1g L 2 Lx + x 2 dx
2 L3
0

E kinetic =

1 M spring L x
=

v1g dx
2 L 0 L

E kinetic =
E kinetic

1 M spring
x
1 v1g dx
2 0 L L

E kinetic =

4.2.3.2 Effective Mass of a Cantilevered Beam


If a beam acting as a spring also stores a significant amount
of kinetic energy during its deflection, then the lumped parameter model of the beam must include the effective mass
of the beam. We will use a cantilevered beam with a constant
cross-section and properties for this example, Fig.4.9.
We derived the deflected shape of a cantilevered beam
with uniform cross-section loaded at its end, Eq.4.12, and
the spring constant K, Eq.4.13, in Sect.4.2.2.4

(4.20)

V ( x)
d 3 y ( x)

=0
EI
dx 3

4.2 Modeling Translational Elements

207

0.0
-0.2

y(x)
___
y(L)

The contribution of the mass element dm to the kinetic energy of the beam is:

Mode 1 shape,
Eq. 4.21

-0.4
-0.6

0.0

0.2

0.4

0.6

x
_
L

0.8

1.0

x
(4.21)
y1 ( x ) = ymax 1 cos
2L

where ymax=y(L).
The deflected shape calculated from beam theory, used to
determine the elastic strain energy of the beam, and the deflected shape assumed by vibration theory to calculate the kinetic energy of the beam differ slightly as per Fig.4.10. The
deflection in the middle of the cantilever is slightly greater
under the end load than in first mode shape. If we were to assume that the beam oscillated freely in its end-loaded shape,
we would calculate a larger effective mass than that calculated assuming that beam vibrates in the first mode shape.
The first mode stores most of the mechanical energy of a vibration. The harmonics, or frequencies of the higher modes,
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
is called a standing wave, because the sinusoidal oscillation
does not progress down the beam. The amplitude of the oscillation varies cyclically, but the locations of the maxima and
the nodes of zero deflection do not change location.
Since the end of the cantilevered beam at x=L is not loaded in a free vibration, the moment and shear at x=L are zero:
d 2 y ( L)

M ( L)
=0
EI

d 3 y ( L)
dx

V ( L)
=0
EI

The velocity, v(x), of a mass element of the beam, dm, at distance x along the beam is the time derivative of the deflection
y(x) at that location


v ( x) =

The uniform cross-section and properties of the beam allow


the mass element dm to be expressed as the mass per unit
length of the beam


The first mode shape of a cantilevered beam is quarter cycle


of a cosine,

dx

dy ( x )
dt

(4.23)

Fig. 4.10 Normalized plots of the deflected shape of a cantilevered


beam due to an end load, Eq.4.12, and the first mode shape of a vibrating cantilever, Eq.4.25

dy ( x )
1
d E M ( x ) = dm
2
dt

End load bending


displacement, Eq. 4.12

-0.8
-1.0

1
2
d E M ( x ) = dm v ( x )
2

(4.22)

dm = A =

M beam
L

(4.24)

The kinetic energy of the mass element at distance x from


the support is:
1 M beam dy ( x )
(4.25)
d EM ( x) =
2 L dt
2

We assume that the velocity dy(x)/dt varies along the beam


velocity with the same function as displacement, as we did to
calculate the effective mass of a spring, with the maximum
velocity at the end, vmax = v( L) v1g .


x
v ( x ) = vmax 1 cos
2L

x
v ( x ) = v1g 1 cos
2L

(4.26)

Substituting into the expression for the kinetic energy of a


mass element
d EM ( x) =

1 M beam
2 L


x
v1g 1 cos 2 L

Integrating yields the kinetic energy of the beam


L

1 M beam
0 d E M ( x ) = 0 2 L

EM =

EM =


x
v1g 1 cos 2 L dx
L

1 M beam 2
x
v1g 1 cos dx
2L

2 L
0
L

1 M beam 2
x
x
v1g 1 2 cos + cos dx

2L
2 L
2
L
0

1 M beam 2
4L
x x L
x
v1g x
sin + +
sin
EM =
2 L 2 2
L

2 L

4 Mechanical Systems

208

1 M beam 2 3L 4 L
L
L
L
v1g
sin +

EM =
sin
L
2L

2 L
2
2
1
0

EM =

-5

-2

-1

m
v12 , ______
sec

-3
-4

Fig. 4.11 Linear translational damping. The slope of the line is the
damping coefficient b

1
1
M beam v12g = E M = M beam v12g 0.227
2 effective
2
= 0.227 M beam

-3

-2

We now equate the kinetic energy of the beam to the equation for kinetic energy written in terms of the effective mass
of the beam and the velocity of the end at x=L

effective

-4

-1

1
M beam v12g 0.227
2

M beam

1
3 8
M beam v12g
2
2

EM =

1 M beam 2 3 L 2 4 L
v1g

2 L
2 2

EM =

Fb, kN

(4.27)

4.2.4Damper: Viscous Friction Energy


Dissipation
Energy is dissipated in a mechanical system by friction as
heat. Friction is the result of shear, either the shear between
two solid surfaces, Coulomb friction, the shear of a fluid,
or the plastic deformation of a solid. The frictional properties of mechanical systems generally introduce the greatest uncertainty and error in energetic models. The type of
friction that can be reasonably accurately represented in a
dynamic model is viscous friction, which was modeled in
Sect.3.2.1. Viscous friction describes the relationship between shear force and translational shear velocity created by
a thin lubricant film between solid surfaces. Under the proper
combination of geometry, lubricant, and relative velocity between the two surfaces, the shear force required to move one
surface parallel to the other is approximately proportional to
the shear rate, where the shear rate is equal to the difference
in velocity of the two surfaces divided by the distance between them. Ideal viscous friction has a linear relationship
between shear force and shear rate. We apply the model of
ideal viscous friction to dashpots and other damping devices,
in which energy is dissipated in the flow of a viscous fluid
through orifices, tubes, and filters, by formulating a linear relationship between the power variables associated the shear
or pumping a fluid. In a translational mechanical system, the
power dissipated in ideal viscous friction is the product of
the force acting through the damper and the velocity difference or drop across the damper, Eq.3.11 and Fig.4.11.

Fb = bv1g

(3.11)

Different symbols are used in different subdisciplines to represent the same physical quantity. In the area of mechanical
vibrations, the symbol c represents a linear viscous damping coefficient b. Likewise, linear viscous damping is known
as proportional damping in vibrations. In viscoelasticity,
Sect. 4.4.6, is the symbol of a linear viscous damping
coefficient.

4.2.4.1Linearization
We now investigate how to create a linearized model of the
typically non-linear relationship between damper force, Fb,
and the velocity difference across the damper. The behavior of all real systems is non-linear to some degree. Recall
the simple test for linearity from Sect.2.3, Does doubling
the input, double the output? Eq.3.11 is a linear relationship. Doubling the difference in velocity between the piston
and cylinder, v12, doubles the force, Fb, acting through the
damper.
The elemental equation

Fb = b1v12
(4.28)
is linear. However,

Fb = b1v12 + b2 v12 v12 + b3 v123

(4.29)

Fb = b1v12 + b2 v12 v12


(4.30)
Fb = b2 v12 v12
(4.31)
are non-linear relationships; doubling v12 does not double Fb.
Note that the squared term is expressed as the product of v12
and its absolute value, rather than as v122 to preserve the effect
of the sign of the velocity on the resulting force.
If the operating range of v12 is centered on zero, then we
linearize, or approximate with a linear equation, a relation-

4.2 Modeling Translational Elements

209

Fig. 4.12a Non-linear viscous


friction. b Linearized models of
viscous friction

Damper
Force Fb

Damper
Force Fb

b II
bI

Velocity, v12

Velocity, v12
Linear Viscous
Friction Models

1,000

3
Fb(v12 ) = 10 v12+ |v12 |v12+ v12

500

Fb, N
Damper
Force

Fb(v12) 37 v12

-500
-1,000
-10

-5

m
v12 Velocity Difference, ___
sec

10

which may or may not be centered in the middle of the operating range. Rather than using a formal Taylor series where
the coefficient of the second term is the derivative of the function at the operating point, and the straight line is tangent to
the curve at the operating point, we may choose to fit a line
to the curve at the operating point, by minimizing either the
absolute value or the square of the error. In any case, the linearization is the sum of the value of the function at the operating point, plus a term which is the product of a coefficient and
the difference between the value v12 and the operating point


Fig. 4.13 Linearization of non-linear translational damping

ship of the form, Eq.4.17, by truncating the squared and


cubed terms and then adjusting magnitude of the coefficient
b1 to best fit the curve over the operating range, Fig. 4.12.
The criterion for best is generally left to the judgment of
the engineer. Fitting a straight line to the curve by minimizing the sum of either the square or the absolute value of the
error between the non-linear and linear relationships is a
common method. For example, approximating
Fb = 10

Fb 37

N sec
v12
m

by a linear relationship to fit the operating range


6 m/sec < v12 < 6 m/sec with the minimum sum of the
error squared requires increasing the coefficient of the firstorder term from 10 to 37; whereas, the fit, which minimized
the sum of the absolute value of the error, yields the coefficient 33, Fig.4.13.
If the operating range of the system is not centered on
zero, then we linearize non-linear relationships with a twoterm Taylor series expansion at the expected operating point,

vop . pt .

+ v12 vop. pt .

dF
) dv

(4.32)

12 vop . pt .

For example, if the non-linear damping relationship


Fb = 50

N sec
v12 v12
m

was to be linearized for the operating point v12 = 2 m/sec and


the operating range from v12 = 1m/sec to v12 = 3 m/sec then
we create the two-term Taylor series expansion.
Fb = 50

N sec
N sec 2
N sec3 3
v12 + 10
v12 v12 + 1
v12
m
m
m

as

Fb = Fb

N sec 2 m
m
2
2
sec sec
m2
N sec 2 m
m

2
+ v12 2
(
)

50 m 2 2 sec

sec

m
N sec

Fb = 200 N + v12 2
200

sec
m

(4.33)

4.2.4.2 Shock Absorbers and Snubbers


Shock absorbers and snubbers (large shock absorbers)
are dampers used to dissipate the energy of disturbance forces and motions. They have a piston-cylinder design and are
the basis for the symbol of a dashpot. There are a variety
of designs which differ in how the working fluid, oil, flows
from one side of the piston to the other under load. The most
common example are the shock absorbers used as dampers

4 Mechanical Systems

210
900

pressure tube into the outer reserve tube. The two valves
have multiple valve stages, which are disks with orifices
that become exposed and pass flow at different pressures.
The disk stack presses against a compression spring which
controls the valve opening.
The forcevelocity difference traces of Fig.4.15 are
clearly non-linear. A less obviously non-linear relationship is

Fb(v12 ) = 50 |v12|v12

500

F b, N

Damper
Force

dFb(2)
m
___
Fb(v12 ) Fb(2)+ ________ v12 - 2 sec
dv12

-500
-900

-4

-2

m
v12 Velocity Difference, ___
sec

Fig. 4.14 Linearization of non-linear translational damping about a


non-zero operating point

500

Fb = bv12 + Fb0

where Fb0 is constant

Although it is the equation for a straight line with familiar


form, y = mx + b, the line does not pass through the origin.
It is offset vertically by the additive constant. Doubling v12
does not double Fb. A straight line that does not pass through
the origin does not describe a linear model. It is an incrementally linear model. Compare the plot of Fig.4.16a and b.
The linear model is described by Fb = bv12, in which the
force is zero at zero velocity. The incrementally linear model
is described by Fb = bv12 + Fb . The offset Fb is the non-zero
force at zero velocity. The test does double the input yield
double the output fails. Superposition cannot be applied.
Any functional relationship which does not pass through the
origin is non-linear.
Linear ordinary differential equations with constant coefficients result from using linear elemental equations with
constant elemental parameters to model an energetic system.
Linear elemental equations are often only poor approximations of the behavior of components in real systems. Common
machine elements with significantly non-linear behavior are
energy dissipaters, where Coulomb friction is particularly
troublesome, and transformers and transducers are subject
to saturation and power sources with power limits. Unfortunately, the only type of differential equation of practical
importance we can solve analytically is a linear ordinary differential equation with constant coefficients. Consequently,
we must use linear elemental equations, even when we know
they are only poor approximation, because our design techniques require linear differential equations.
When we need to improve the accuracy of our mathematical model due to the significance of the systems nonlinear behavior, we use numerical solutions, such as the
RungeKutta method, which allow us to include non-linear
elements. We will introduce numerical method in Chap.7.
However, even when we use a numerical solution, we will
still start our analysis with a linear approximation. Ironically,
even when a linearized model is too inaccurate to yield useful results, it still has one useful property: a linear model has
only one solution. Non-linear models can have a number of
solutions. If we use the inaccurate results of our linear model
as a starting point to begin the iteration required to fit a nonlinear model to our data, we save time and are more likely to
converge to a physically meaningful solution.
0

Fb
lbf

250

-250
-10

-5

in
v12 , ___
sec

10

Fig. 4.15 Shock absorber forcevelocity difference curves in US customary units

in vehicle suspensions. A shock mount has no working


fluid. It is an elastomer, a rubbery polymer or synthetic rubber. The elastomeric motor mounts between the engine of
a vehicle and its frame or subframe to reduce vibration are
shock mounts.
A shock absorber or snubber with a fixed orifice diameter has a forcevelocity relationship similar to the trace in
Fig. 4.14, where the damper becomes stiffer with increasing velocity difference between the piston and cylinder. This
response is unacceptable for a vehicle suspension. Vehicle
shock absorbers are designed to produce either an approximately linear or a digressive forcevelocity difference
curve, where digressive refers to a curve which becomes softer (flatter) with increasing velocity difference, per Fig.4.15.
The forcevelocity difference relationship is created by the
valve design. Conventional automotive shock absorbers
have two valves, one in the piston, which controls flow of oil
from one side of the piston to the other, and a second valve at
the base of the cylinder, which controls flow from the inner

(4.34)

4.2 Modeling Translational Elements


Fig. 4.16a Linear model.
b Incrementally linear model.
The equation of a straight line
which does not intersect the
origin is incrementally linear

211

Damper
Force Fb

Damper
Force Fb

Velocity, v12

Velocity, v12

Line passes
through the origin

Fig. 4.17a Static and kinetic


friction coefficients model of
Coulomb friction. b Typical
average actual friction force
plotted with the friction coefficients model. The friction force
has the opposite sign of the force
acting through the damper due to
inverse sign conventions

a
Kinetic Friction

Friction
Force

Line does not pass


through the origin

Average Actual
Friction Force

Friction
Force

Static Friction
Velocity

Consider the static and kinetic model of Coulomb friction, described in terms of coefficients of friction. The model
consists of two horizontal lines for v 0 and two points for
v=0, Fig. 4.17a. It yields four different values of friction,
two for the static case and two for the kinetic case, which
depend on the direction of motion, since the frictional force
opposes motion. Typical measured Coulomb friction is more
difficult to describe, Fig.4.17b.
In the case of Coulomb friction, the conditions of the surfaces in contact are subject to change. Surface contaminants,
i.e., dirt and wear particles, and chemical and mechanical alteration of a surface change its frictional properties. To start
with, there are many different types of friction. The most
familiar is dry friction between two surfaces. A very useful model consists of static and kinetic friction coefficients,
which are presented as constants. Although this model is
very useful, it isnt very realistic. Shouldnt the relationship
between the normal and tangential force acting on surfaces
depend on the displacement of the surfaces, not to mention
the wear, temperature, lubrication, cleanliness, and chemical alteration of the surfaces? Friction depends on all these
factors. Further, there is a stochastic, or probabilistic, aspect
evident in the transition from kinetic to static friction; or, in
other words, when a sliding object stops.
Although the static and kinetic Coulomb friction model
is conceptually simple, it is not mathematically simple,
because it is non-linear and consists of two lines (and two

Velocity

points), rather than one straight line. We must simplify the


model to use it analytically. The first thing we will do is neglect the case of static friction, since we are developing a
dynamic model. If a mechanical system isnt moving, then
it isnt dynamic. The response either hasnt started yet, or it
has reached steady-state. Further, it is difficult to implement
mathematically. How is our model to know which way to
apply a static force to oppose motion which hasnt happened
yet? We can model kinetic friction, if we know the direction
of motion, by using a force source to apply a constant force
to represent the kinetic friction resisting motion. This method works well for first-order dynamic systems, since they
cannot oscillate, and we will know the direction of the motion. The only trouble is that we cannot turn off the kinetic
friction when the velocity reaches zero. We must keep this in
mind when interpreting the results of our model.

4.2.5 Translational Mechanical Sources


Although some energetic mechanical systems operate from
energy stored in springs or flywheel over the period we wish
to model them, it is more common that the system draws energy from a power supply. Mechanical power is the product
of force and velocity, P = Fv. Each element in a linear graph
represents a power flow in the energetic system, where the
power sourced, stored, or dissipated by the element equals

4 Mechanical Systems

212

x,v

F(t)

b
M

F(t)

g
Fig. 4.18 Schematic and linear graph of a force source-mass-damper
system. The power flow in each element is the product of the through
variable, force, and the across variable, velocity

the product of the force acting through the element and the
velocity difference across the element.
A power source is identified by which of its two power variables of the system is either known or controlled. Mechanical
systems have force sources and velocity sources. If a source is
powerful enough, it is possible to control (or specify) the time
history of one, but not both, of its two power variables. The
magnitude of the second power variable is determined by the
response of the system. For example, if power is supplied to
a system at a specified force, then the response of the system
determines the velocity of application of that force.
An abstract force source, Fig.4.18, is the mechanical
power source most familiar to engineering students, since
they have seen many vectors labeled F or F(t) on free body
diagrams. The convention of representing power sources as
vectors indicates that whatever is providing power to the
system is powerful enough that it can maintain the specified
value of the input variable. In order to apply the specified
force to the system, the point of application of the force must
move with the velocity of the system. If the force source
were to lose contact, then it could not apply the force. Consequently, a force source must also be capable of displacement at the velocity of the system it is driving. A mechanical
power source cannot float in space and apply force to an
object, as the input force vector is portrayed in Fig.4.18. A
mechanical source must react against something, since for
action there must be reaction, per Fig.4.19a. If a force vector
is shown floating in space, then the necessary reaction 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
able to supply enough power such that it can approximately
maintain the specified magnitude of the input variable, force
Fig. 4.19a Force source massdamper system shown with
support to provide the reaction
needed by the mechanical power
source. b Conventional representation of the power source as a
vector with the reaction against
ground implied

a
g

or velocity, as the system draws power in response to the


change of the input. Force sources must be powerful enough
to maintain the specified force over the range of the velocity
of the point of application of the force. Likewise, velocity
sources must be powerful enough to maintain the specified
velocity, by applying the force needed to move the point of
application of the force at the specified velocity. The more
powerful the device, the more nearly perfectly it can behave
as a source. On earth, gravity is a nearly perfect source of
force acting on a mass, since the gravitational force is virtually 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
source, depending on how they are controlled. For example,
laboratory tensile testing machines typically can be placed
under either force or velocity control. A powerful hydraulic
piston under feedback control can function as either a force
source or a velocity source. Again, if we control the force
applied by the source, we cannot control the velocity of the
point of application, unless we dictate the laws of physics.
Similarly, a velocity source for a device provides power to
a mechanical system with a known or prescribed velocity. If
we control the velocity of the point of application of a velocity source, we cannot control the force that the source applies.
Force sources, illustrated as force vectors, are the typical
mechanical source used in introductory courses. However,
in machine design, energy is often provided by a velocity
source. A velocity source is a mechanical element that follows
a specified motion at a specified velocity. Whatever is moving the machine element must be powerful enough to provide
the force necessary to move the point of application of the
source at the specified velocity. Velocity sources are more
familiar than they seem at first. We more commonly set or
control the speed of a machine than we do the force it applies.
Power limitation determines what type of source can
act on what type of element. We will classify both sources
and energy storage elements by one of the power variables,
force or velocity in mechanical systems. We can control only
one of the two power variables of a power source. As stated
above, a force source is a device which provides power to a
mechanical system with a known or prescribed force.

4.2.5.1 Limitations of Mechanical Sources


The limitations of mechanical sources are imposed by the magnitude of the power drawn by the system. Ideal power sources

x,v

F(t)

1
M

b
g g

x,v

F(t)

1
M

b
g

4.2 Modeling Translational Elements

213

The corresponding limitation applies to strain energy. We


cannot apply a step change in force to a spring with a force
source, since it too would require infinite power. From the
perspective of power and energy,

x,v

v(t)

v(t)

lim

t 0

g
Fig. 4.20 Schematic and linear graph of a velocity source-massdamper system. It is impossible to command a step change in velocity
because that would require infinite power from the source

maintain the specified value of the input variable regardless


of the power drawn by the system. However, for an energetic
model to be useful for all possible inputs which might be applied to the system, a power source cannot directly drive an
energy storage element with the state (energy storage) variable
identical to the source.
The two mechanical system energy storage elements, a
mass and a spring, store kinetic energy and strain energy, in
terms of velocity and force, respectively

Em =

1 2
mv1g
2

EK =

F2
1
1 F
Kx12 2 = K K = K
2
2 K
2K

If we control a source, then we must be able to command it


to apply any arbitrary input to the system. A Heaviside step
input is an extreme input, since it instantaneously changes values. We cannot apply a step change in velocity to a
mass with a velocity source, Fig.4.20. A finite change in
the amount of energy stored in the mass over the infinitesimal time period from t = 0 to t = 0+ would require infinite
power from the source
lim

t 0

E M E M E M
=

P =
dt
0
t

From the Newtonian perspective, a finite change in velocity over infinitesimal time is an infinite acceleration, which
would require an infinite force to impose on a mass.

Fig. 4.21a Schematic and


b linear graph of a force source
spring-damper system. It is
impossible to command a step
change in velocity because that
would require infinite power
from the source

E K E K E K
=

P =
dt
0
t

From the Newtonian perspective, it is impossible to apply


a Heaviside step input of force directly to a spring, since
the point of application of the force must displace to create
the spring force, Fig. 4.21. Instantaneous deformation of a
spring requires infinite velocity, which is impossible.
The restriction that a force source cannot impose a force
on a spring runs contrary to many models you have used in
statics and machine design. A force represented as a vector
acting directly on a spring is a common model in physics
and engineering texts. What is unstated in those models is
that either (1) the system has reached steady-state, and the
force was not constant during the transient period, or (2)
the input function increased the applied force gradually, so
that the power drawn from the source by the system did
not exceed the capacity of the source. We will see when
we investigate frequency response in Chap.10 that a force
source can act directly on a spring, because a sinusoidal
input varies gradually enough to limit the power demanded
from the source. Force inputs which start at zero then ramp
up or down can also be applied directly to springs. Textbook figures showing force sources acting on springs imply
one of these restrictions or special cases, because in the
general case, when there are no restrictions on the type of
input function, a force source cannot arbitrarily control the
amount of energy stored in a spring.
Violation of the restriction that a force source cannot impose a force directly on a spring is more difficult to detect,
because we need to look beyond the element which is the
point of application. For example, a force source acting on
a damper seems reasonable, until it is recognized that the
damper is in series spring, and the two elements carry the
same force.

x,v

F(t)

b
b

F(t)
g

4 Mechanical Systems

214
Fig. 4.22a Schematic and
b linear graph of a force source
damper-spring system. Although
the force source F(t) acts on
damper b, it also imposes its
force on spring K. It is impossible to apply a Heaviside step
input to this system

b
x,v

F(t)

F(t)
g

Fig. 4.23a Schematic and b


linear graph of a force source
damper-spring system with a
mass element added at the point
of application of the force source.
The force applied by the source
divides at node 1 between accelerating mass M and compressing
damper b in series with spring K.
At the instant after a step change
in force, all of the step change
acts to accelerate mass M

x,v

F(t)

b
M

The concern of demanding infinite power from a source


commanded to undergo the instantaneous transition of a
Heaviside step may seem artificial, since no physically real
power source can instantaneously change the magnitude
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
using a model in which we attempt to dictate the value of an
energy storage (state) variable is that the system cannot have
a transient response. This is why models of that sort are only
in steady-state.
If we were modeling a physical system with the dominate
energetic properties of strain energy storage and energy dissipation; and the input to the system was better modeled as
a force source than a velocity source, we would modify the
model shown in Fig.4.22 by adding a mass element at the
point of application of the source, per Fig.4.23. The amount
of energy stored in the mass element is small enough, so that
we would usually neglect it. However, in this case, it serves
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 deformation of the spring, thereby limiting the power flow
from the source.
A significant problem for modeling energetic systems is
that linear systems must have a source which both provides
(source) and accepts (sink) power, while maintaining the specified value of one of its two power variables. There are relatively few sources which can push and pull with the same force
or velocity. We must ignore this limitation of the hardware

F(t)

M
g

in order to use linear models. If it is the case that the source


can only push and not pull, or only move forward but not in
reverse, then we are limited to modeling only part of the response of the system without resorting to numerical solutions.

4.2.6Summary
It is important to keep clearly in mind that a schematic is a
model of an energetic system. The energetic properties represented by the elements in the schematic are not all of the properties of the system. They are the dominant energetic properties that must be included, in order to model the behavior
of the system with the precision needed for the particular engineering analysis being performed. You will discover as we
work with dynamic models that what is significant and must
be included in a model for it to be reasonably representative of
the system depends on many factors, including the magnitude
and type of input and the time scale of the response.

4.3 The Sign Problem of Mechanical Systems


We need to consider signs. There are two incompatible definitions of a positive force in mechanical engineering. The
convention we choose to use depends on what is important
to us in the particular analysis. The convention which arose
from the mechanics of rigid bodies defines a force as positive based on the direction of force in space. A positive force
produces a positive acceleration on a mass, that is, an accel-

4.3 The Sign Problem of Mechanical Systems


Fig. 4.24a Force in the positive
direction puts the spring into
tension. b Force in the positive
direction puts the spring into
compression

215

x,v

F(t)

x,v

F(t)

Frictionless rollers

x 1, v 1

x 2 , v2

FK

FK

Node 1

Node 2

Fig. 4.25 Translational spring showing both displacement and velocity variables at the nodes. The force FK is shown in the positive direction at either end of the spring

a
1

b
2

Fig. 4.26a Linear graph spring element oriented by the arrowhead


with the positive direction from node 1 and node 2. b Spring element
oriented from node 2 to node 1

eration in the positive direction as defined by the axes. The


convention which arose from solid or continuum mechanics
defines a positive force based upon its effect on materials
or machine elements. A force which produces tension on an
infinitesimal element is defined as positive.
As it happens, the positive definition of force from continuum mechanics corresponds with the effect of the forces
on the design of machine and structural components. The
common meaning of the word, positive, is good or favorable. For example, we welcome a positive review of our
performance. Tensile force is positive, in the sense of being
favorable, because engineers generally prefer to have elements loaded in tension rather than compression. Compression can lead to buckling, a sudden and catastrophic failure
mode. Tension cannot produce buckling. Elements designed
to carry a compressive are more complicated and more expensive than elements that carry a tensile load of the same
magnitude. Hence, compression is bad, or at least unwanted,
and negative.
Consider the two similar spring-mass systems in Fig.4.24.
They are both acted on by a force F(t) in the positive direction. In Fig.4.24a, positive displacement of mass M places
spring K in tension whereas, in Fig.4.24b, positive displacement of mass M places spring K in compression.
Isolating a translational spring and showing the equal
forces acting on it in the positive direction, Fig.4.25, gets

Frictionless rollers

to the crux of the sign problem. It looks wrong and, indeed,


if equal forces acted in the same direction on both ends of a
translational spring, the spring would not deform.
In a conventional Newtonian analysis, the conflict between the sign conventions based on the direction of the
force, used for mass, and the sign conventions based on the
effect of the force, used for stress, strain, and springs, is resolved by drawing free body diagrams of every mechanical
element in the system. This is a lot of work and, unfortunately, error-prone due to confusing action and reaction forces.
The linear graph sign conventions limit the effort and
possibility of error, by arbitrarily defining a positive direction for the force in acting through each element in the
network, by drawing an arrowhead near the middle of the
line connecting thetwo nodes. This process is known as
orienting the element. The direction of the arrowhead
is arbitrary, but once chosen, it defines the positive direction for the force acting through the element and, therefore,
the order of the nodes for calculating the positive velocity
difference across the element. The direction a force acts
on a node will be interpreted using a flow analogy and the
bank account sign convention. Force directed into a node
is positive. Force directed out of a node is negative. The
arbitrary positive directions of elements work out the same
way that arbitrary signs assumed in a truss analysis are corrected by the algebra.
For example, the positive direction for the spring in
Fig. 4.26a is from node 1 to node 2, yielding the element
equation
dFK
= Kv12
dt
Conversely, the positive direction for the spring in Fig.4.26b
is defined by the arrowhead as from node 2 to node 1. Consequently, the element equation is
dFK
= Kv21
dt
Recall that reversing the order of the node subscripts on velocity differences is equivalent to inverting the sign,
v12 = v21

4 Mechanical Systems

216

Fig. 4.27a Linear graph of a


spring-mass-damper system with
the positive direction damper
band 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 directions of damper b and spring
K reversed

F(t)

F(t)

g
We will demonstrate that reversing the arbitrarily defined
positive directions for elements does not affect the resulting
differential system equation, by reducing the linear graphs in
Fig.4.27a and b for the velocity of the mass, v1g. The linear
graphs have opposite directions defined as positive for damper b and spring K. The sign reversals of the through variables,
forces Fb and FK, are included in the continuity and element
equations. The sign reversals of the velocity differences, v12
and v2g, appear in the compatibility and element equations.
The continuity, compatibility, and element equations of
the linear graph in Fig.4.27a are
Continuity:

F (t ) = FM + Fb

Compatibility:
Elements:

Fb = FK

v1g = v12 + v2 g

Fb = bv12

FM = M

dv1g

dFK
= Kv12
dt

dt

Reducing these equations to the system equation for the


velocity of the mass, v1g
F (t ) = FM + Fb

F (t ) = M

dv1g

F (t ) = M

dt

F (t ) = M

dv1g
dt

dv1g
dt

+ b v1g v2 g

+ bv12

+ bv1g bv2 g

dv1g

b dFK
+ bv1g
F (t ) = M
dt
K dt
F (t ) = M
F (t ) = M

dv1g
dt

dv1g
dt

+ bv1g

+ bv1g

g
2
dv1g
b dF (t )
bM d v1g
+ F (t ) = M
+ bv1g +
K dt
dt
K dt 2

d 2 v1g K dv1g K
1 dF (t ) K
+
F (t ) =
+
+ v1g
M dt
bM
b dt
M
dt 2
The equations for the linear graph in Fig.4.27b, in which the
positive direction for damper b and spring K are the reverse
of the linear graph in Fig.4.27a are
Continuity:

F (t ) + Fb = FM

Compatibility:

v1g = v1g

b dFb
K dt

b d
( F (t ) FM )
K dt

dv1g
b dF (t )
b dFM
+ F (t ) = M
+ bv1g +
K dt
dt
K dt

Elements:

Fb = FK

v1g = v21 vg 2

Fb = bv21

FM = M

v1g = v1g
dv1g

dFK
= Kv21
dt

dt

Reducing the second set of energetic equations to the system


equation for the velocity of the mass, v1g, will yield the previous result
F (t ) + Fb = FM

F (t ) = Fb + FM

F (t ) = bv21 + M

dv1g
dt

F (t ) = b v1g vg 2 + M
F (t ) = bv1g + bvg 2 + M
F (t ) = bv1g +
F (t ) = bv1g +

dv1g
dt

dv1g
dt

dv1g
b dFb
+M
K dt
dt

dv1g
b d
F (t ) + FM ) + M
(
K dt
dt

dv1g
b dFM
b dF (t )
+ F (t ) = bv1g +
+M
K dt
K dt
dt
2
dv1g
bM d v1g
b dF (t )
+ F (t ) = bv1g +
+M
2
K dt
dt
K dt

4.4 Drawing Linear Graphs from Mechanical Schematics


Fig. 4.28 Linear graph symbols
for translational mechanical
systems

217

F(t)

v(t)

g
Force Source
d 2 v1g K dv1g K
1 dF (t ) K
+
F (t ) =
+
+ v1g
M dt
bM
b dt
M
dt 2
As long as we are consistent in using the arbitrarily defined
positive direction for each element in the linear graph, the
signs of the equations will be consistent. The convention of
declaring an arbitrary positive direction for each element
simplifies the reduction of the linear graph networks to a
differential system equation. It is at the end, after we have
a system equation, that we will use a free body diagram to
establish the relationships between the positive directions for
the input and output variables in the linear graph with the
forces and velocities in the actual physical system.
We will always need to exercise care in interpreting and
communicating the results of our analyses. Graphical definition of the positive directions of the input and output variables on the mechanical schematic (physical model) of the
system is the most effective method.
We shall see that the different sign conventions of subsystems of dissimilar types or energy domains may be
contradictory. Unfortunately, the signs of systems with
energy transformation or conversion are further complicated by the way subsystems are coupled in a linear graph.
Further, when a differential system equation is used as the
model of a plant placed under feedback control, its sign
must be fixed, so that a positive input to the system yields
a positive value of the output variable or the closed-loop
feedback control system will be unstable. For the meantime, we will accept the sign yielded by the linear graph
method and not trouble ourselves to determine how that
sign corresponds to one of the many sign conventions used
in mechanical engineering.

4.4Drawing Linear Graphs from Mechanical


Schematics
The first step in drawing a linear graph is to find the nodes of
distinct values of the across variable and identify them on the
schematic of the system. Once you have identified the nodes
on the schematic, draw the linear graph by first drawing and

g
Velocity Source

Mass

Spring

Damper

labeling the nodes and then adding the elements between


their respective nodes.

4.4.1 Linear Graph Symbols


The across variable in mechanical systems is velocity.
Springs and dampers have a node at both ends, since they
extend and compress. However, an ideal mass has a single
velocity because it is rigid. The other velocity node associated with the mass is the velocity of the ground node. The
linear graph element is drawn between the node representing the velocity of the mass and the ground node. Ground
does not supply a reaction force for a mass, only an inertial
velocity reference. All of the force flowing into a mass element acts to accelerate the mass. The lower half of a branch
representing a mass is dashed to indicate that the force is not
transmitted to ground.
Each branch of a linear graph represents a power flow.
Power can flow into or out of the energy storage and source
elements. However, power can only flow into damper
elements where it is dissipated as heat and lost from the mechanical system. Power sources differ from passive elements. Sources pump energy uphill; the across variable
velocity increases in the positive direction of the through
variable force in a source. In all other elements, the across
variable, decreases in the positive direction of the through
variable. Source symbols also differ from the symbols of
the other elements. The convention from electrical circuits
is used, and sources are identified by a circle containing the
power variable under our control, Fig.4.28.

4.4.2Force Source Acting on a Parallel


Mass-Damper System
The ideal, massless and rigid bar shown in Figs. 4.29 and
4.30 is a force spreader which distributes the input force to
mass M and damper b. Since the ideal bar is rigid and massless, it has no energetic properties. It functions as an ideal
force conductor. The bar cannot rotate. This constraint is

4 Mechanical Systems

218
Fig. 4.29a An energetic
schematic of a force source
mass-damper system. b Nodes of
distinct values of the across variable velocity located and labeled
on the schematic. The rigid bar
has a single translational velocity. The bar and the ideal rigid
rods connected to have the same
velocity

Ideal massless
Ideal massless
and rigid rods
and rigid bar x,v

b
x,v

1
F(t)

F(t)

g
1
M
g

Fig. 4.30a Redundant nodes 1


removed but redundant ground
nodes retained. b Equivalent
force source-mass-damper
system

x,v

b
g

F(t)

x,v

F(t)

g
Fig. 4.31a The linear graph
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
applied to the system, all of the
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
the damper and is dissipated as
heat; no power flows into the mass

b
M

1
F(t)

Power flow during


transient period

implied, since only the x-axis identified. The horizontal lines


connected to the piston and cylinder of the dashpot and to
the mass are ideal, massless and rigid rods. The velocity difference which can exist across the dashpot is between the
piston and the cylinder. In an actual shock absorber, the deformation of the steel components is minute, relative to the
displacement of the piston within the cylinder, validating this
model of a damper.
Draw the linear graph by drawing and identifying the
across variable nodes and then adding the elements between the appropriate nodes, per Fig.4.31. There are just
two unique velocities and, hence, two nodes, node 1 and
ground. We may use multiple ground nodes in large linear
graphs for clarity but never multiples of any other nodes.
The linear graph shows how both the through variables
and power flows divide. The force applied to node 1 by the
force source divides between the mass and the damper, in

1
F(t)

1
M

Power flow at
time t = 0 +

1
F(t)

g
Power flow in
steady-state

the general case. Assume the system is initially de-energized, when a step input of force F(t)=F0us(t) is applied
to the system. From the Newtonian perspective, the system
tends towards force equilibrium. Initially, all of the force
acts to accelerate the mass. When the mass begins to move,
the damper compresses and begins to exert force. Eventually, as the velocity difference across the damper increases,
the force acting to compress the damper equals the applied
force, leaving none to accelerate the mass.
From the complementary energetic perspective, the system tends towards energy equilibrium. When the energy of
the power supply is made available to the de-energized system, the power flow travels into the energy storage element.
As the power flow into the system continues, and the system
approaches an energetic equilibrium with the power supply,
less mechanical energy flows into the kinetic energy storage
element, and more is dissipated as heat by the damper. When

4.4 Drawing Linear Graphs from Mechanical Schematics

Fig. 4.32a Energetic schematic of a force source mass-two


damper system. b Nodes of
distinct values of the across
variable translational velocity
located and labeled

219

Ideal massless and rigid bar


x,v

x,v

b1

F(t)

b2

b1
g

F(t)

b2

2
M
g

b2

F(t)

4.4.4Force Source Acting on System


of a Mass and Two Dampers

b1

A second configuration of a system with a mass and two


dampers is for the force source to act directly on damper
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. However, 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

the system reaches energetic equilibrium with the power


source, no more energy flows into the kinetic energy storage element. All of the mechanical energy flowing into the
system is dissipated as heat.

4.4.3Force Source Acting on a System


of a Mass and Two Dampers
Rigid elements have a single translational velocity. Masses
are rigid and are always assigned a velocity node. The ideal
massless and rigid bar of Fig.4.32, which distributes the applied force to dampers b1 and b2, also has a distinct velocity and is assigned a node. The force source needs a reaction. Since none is shown, we assume that the source reacts
against ground, Fig. 4.33.

Fig. 4.34a Energetic schematic of a force source mass-two


damper system. b Nodes of
distinct values of the across
variable translational velocity
located and labeled

4.4.5Force Source Acting on a Mass-SpringDamper System


A force source acting on a mass-spring-damper system is
shown in Fig.4.36 and its linear graph in Fig.4.37. Note that
there are two ground nodes in this linear graph. The horizontal 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
easier to read.

4.4.6 Viscoelastic Models


Polymers, i.e., plastics, and elastomers, i.e., synthetic rubbers, have both elastic strain energy storage and damping
properties. Materials which combine these two properties at
room temperature and under low stress are called viscoelastic, where the prefix, visco-, denotes the viscous energy dissipation property and the suffix, -elastic, denotes strain energy
storage. Although a plastic, elastomeric, or naturally occurring polymer part or specimen is a single object, it requires

x,v

F(t)

b1

b2
M

x,v

F(t)
g

b1
M

b2
g

4 Mechanical Systems

220

b1

b2

F(t)

g
Fig. 4.35 Linear graph of the force source-mass-two damper system
shown schematically in Fig.4.34

two elements in a lumped parameter model to represent its


energetic behavior: a damper to represent the energy dissipation, and a spring to represent the strain energy storage.
Viscoelasticity uses different symbols for the material parameters than are used in system dynamics. The symbol used
in viscoelasticity for a spring constant is E. The symbol for
a damping constant is . The terminology of viscoelasticity
also differs slightly from system dynamics. A time constant
is known as a relaxation time and given the symbol T. The
term relaxation time has its origin in tests in which a viscoelastic material is put under tension and held at a certain extended length, while the force on the specimen is measured.
The relaxation time is calculated from the decay of the force.
The use of relaxation time has extended beyond viscoelasticity into other areas of engineering, where models are based
on real exponential decays.
When mechanical elements are connected such that they
act to share a load, they are described as acting in parallel. The viscoelastic model with the spring and damper in
parallel was introduced by Kelvin and is known as the KelvinVoight viscoelastic model. When mechanical elements
are connected such that they carry the same load, they are
described as acting in series. This is the Maxwell viscoelastic model.

4.4.6.1 The KelvinVoight Viscoelastic Model


A single spring and dashpot pair cannot replicate non-linear viscoelasticity. The models shown Figs.4.38 and 4.39
are ganged-up with multiple replicas (an infinite number in
theory) in a chain with different damping and stiffness elements, Figs. 4.40 and 4.41. The spring constant and damping coefficients of each spring and dashpot pair is identified
Fig. 4.36a Schematic of a
force source acting on a massspring-damper system. b Nodes
of distinct values of the across
variable velocity

by a numerical subscript. Hence, the last element, the spring


at infinity, is identified as E, or in system dynamics notation, 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
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 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 precision needed to fit the
data. Implementation occurs with a finite number of springs
and dashpots, per Fig.4.40. The KelvinVoight model has
instantaneous displacement due to the spring K , which is
not paired with a dashpot in parallel. As we have established,
it is impossible to apply a Heaviside step input to a spring.
The instantaneous displacement is the displacement that
occurs during the rapid, but not instantaneous, application of
a fixed dead load.

4.4.6.2 The Maxwell Viscoelastic Model


A viscoelastic model which has the spring and damper in
series was introduced by Maxwell, Figs.4.42 and 4.43. We
have considered this model previously, Fig.4.18, and concluded that it is impossible to drive a spring and dashpot in
series with a force source in the general case. We found that
the force source dictated the value of the state variable, the
force acting through the spring, and, thus, eliminated the dynamic response of the energetic system. Consequently, in the
general case, a single spring and dashpot pair in series must
be driven by a velocity source instead of a force source, to
eliminate the impossible instantaneous deformation of the
spring. We shall see when we investigate frequency response
in Chap.10, that a sinusoidal force is a useful test input for
characterizing viscoelastic materials.
Multiple pairs of spring and dashpots of the Maxwell
model are ganged in parallel with a spring, Kn, to approximate a non-linear viscoelastic response, Figs.4.44 and 4.45.
4.4.6.3 The Zener Viscoelastic Model
The Zener viscoelastic model consists of the first and last
elements of the infinitely ganged Maxwell model, a single
Maxwell model in parallel with a spring, Fig.4.46. This
model is popular enough to be termed the standard linear
model in viscoelasticity. The linear graph is Fig.4.47. The

x,v

F(t)

x,v

F(t)
g

1
M
g

b
g

4.6 Modeling Rotational Elements

F(t)

221

Beware that all angular measures are dimensionless. Although we use a degree symbol of superscripted circle, degrees, radians, cycles, and revolutions have no units. Work
in a rotational mechanical system is the product of torque
and angular displacement . Because angular measures are
dimensionless, the units of torque, [T ] = [ F L ], are the same

as the units of mechanical work, [W ] = [ E ] = [ F L ].

W = E = T 12
(4.36)

g
Fig. 4.37 Linear graph of the force source-mass-spring-damper system shown schematically in Fig.4.36

Zener model is shown being driven by a velocity source for


the general case. The steady-state response of the model of
a sinusoidal force input of varying frequency, termed its frequency response, is also used in viscoelastic material testing.

4.5 Rotational Mechanical System Elements


Rotational mechanical systems are directly analogous to
translational mechanical systems. The power variables in
a rotational mechanical system are torque and rotational
velocity. We must express rotational velocity as angular
velocity in radians/sec. Power in rotational mechanical
system is the product of torque and angular velocity.


P = T 12

(4.35)

The easiest way to introduce the equations for rotational systems is by direct analogy to translational systems, Table4.4.
We can use a one-to-one set of analogies to relate rotational
mechanical systems to translational mechanical systems,
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
constants and damping coefficients in the two types of systems. A subscript R, denoting rotation, can be used to
distinguish K and KR and b and bR, but this is not necessary.
It is clear from the variables used in an equation whether the
motion is rotation or translation.
Fig. 4.38a Energetic schematic
of the KelvinVoight viscoelastic model. b KelvinVoight
viscoelastic model with nodes
of distinct values of the across
variable velocity

Ideal massless
and rigid bar

4.6 Modeling Rotational Elements


Machine design relies on rotational mechanical systems, because there is generally no limit on angular displacement.
Translational displacement of machine elements is almost
always limited. Rotation is an independent motion with additional degrees of freedom of movement and must be accounted for separately from translation. Most engineers
find translational motion and translational mechanics much
easier to conceptualize than the corresponding quantities of
rotational motion. A case in point is momentum. We have
a better intuitive sense of the effects of translational momentum p than we have for rotational momentum H, which
is why gyroscopes are so fascinating. The parameters of
rotational mechanics are also more difficult to estimate than
their translational analogs. We have an intuitive sense of the
acceleration of a mass under the influence of a force. If the
object is of one density, the mass scales directly with size.
The rotational acceleration of inertia due to a torque is not
as intuitive, because both the quantity and the distribution of
mass relative to the axis of rotation are factors in the mass
moment of inertia. The need to use SI units increases the difficulty of estimating rotational parameter.

4.6.1 Mass Moment of Inertia


Recall from your study of dynamics that the relationship for
rotational analogous to F=ma is
T = I

x,v

x,v

K
F(t)

K
F(t)
g

b
g

4 Mechanical Systems

222

F(t)

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.
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
to accelerate the infinitesimal mass is tangential to the circular motion. Its magnitude is

g
Fig. 4.39 Linear graph of the KelvinVoight viscoelastic model shown
in Fig.4.38

where T is a torque acting to accelerate an inertia in rotation,


I is the mass moment of inertia, and is the angular acceleration. We will use this expression, but will restrict rotation to
one axis, so that we can use scalar mathematics, and we will
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
use I for the area moment of inertia in bending or for electric
current. For brevity, we will also refer to the mass moment
of inertia J as rotational inertia.
The rotational equivalent of mass M is mass moment of inertia J. The inertial effect of mass in resisting acceleration in
rotation depends on both the amount of mass and the square
of its distance from the axis of rotation. An ideal mass moment of inertia J is rigid, because its only energetic property
is inertia. If it were able to deform, then it could dissipate energy due to internal or external friction or store strain energy.
We will write the elemental equation for a mass moment
of inertia J, T=J, with angular acceleration expressed as
the first derivative angular velocity , so that the equation is
in terms of the power variables of a rotational system,


TJ = J

d 1g

(4.37)

dt

The angular velocity of rotational inertia must be referenced


to ground!
The mechanics of rotation can be understood in terms of
translational motion. If we examine a mass moment of inertia taking an infinitesimal mass dm, such that the dimen-

Fig. 4.40 Energetic schematic


of the KelvinVoight viscoelastic
model with n-1 dashpots and n
springs

dT
= dFm
r
The instantaneous tangential velocity is the product of the
angular velocity W and the radius r from the axis of rotation
vtangental = r
Writing F = Ma in terms of the applied torque and the angular velocity
dFm = dm

dvtangental

dt

dT
d r
d
= dm
dT = dmr 2
r
dt
dt

All infinitesimal elements of mass dm at radius r from the


axis of rotation must undergo the same acceleration under
the same infinitesimal torque dT, since they must rotate together in a rigid body. Integrating over all infinitesimal mass
elements dm at radius r yields the torque T and the mass of
the mass moment of inertia,

dT = r

d
d
dm T = Mr 2

dt
dt

The inertial effect of the mass at radius r felt at the axis of


rotation is the mass m
oment of inertia
Mass Moment of Inertia J = Mr 2
It is generally the case in machine design that a rotating
mass cannot be reasonably modeled as a point mass or a thin
walled cylinder or ring to keep the mass or the rotational

x,v
K1
F(t)

K2

K n-1
Kn

b1

b2

b n-1

4.6 Modeling Rotational Elements

223

Fig. 4.41 Linear graph of


the KelvinVoight model of
Fig.4.40

F(t)

K1

K2

Kn-1

b1

b2

bn-1

Kn
g

Fig. 4.42a Energetic schematic


of the Maxwell viscoelastic
model. b Maxwell viscoelastic
model with nodes of the across
variable translational velocity

x,v

v(t)

x,v
v(t)

b
g

v(t)

extension of our previous analysis modified, so that the


mass of an infinitesimal ring is a function of its distance
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 infinitesimal volume formulated in terms of the distance r from the
axis of rotation
Mass Moment of Inertia J =

g
Fig. 4.43 Linear graph of the Maxwell viscoelastic model shown in
Fig.4.42

inertia a radius r from the axis. Although the mass of the part
at the greatest radius from the axis of rotation has the greatest contribution to the mass moment of inertia, the rotational
inertia of mass closer to the axis is typically not negligible.
In these cases, the mass moment of inertia must either be
calculated using integral calculus or with tables.
Simplest integral formulation of mass moment of inertia is that for a revolved body. The formulation is an

r dM = r d vol
2

mass

vol

For example, the mass moment of inertia of a cylinder rotating about its axis is formulated by representing the infinitesimal mass as a ring with circumference 2 r , thickness dr, and
length of the cylinder, L, Fig.4.49,
R

J = L 2 r 2 r dr J = L 2
0

r4
4

J = L
0

R4
2

When M is substituted for the product of the materials density and the volume of the cylinder, we find that the mass
moment of inertia of a uniform cylinder rotating about its

Fig. 4.44 Schematic of the


Maxwell viscoelastic model
Ideal massless and rigid bar

F(t)
x,v

K1

K2

Kn-1
Kn

b1

b2

bn-1

4 Mechanical Systems

224

b1 b2
2

F(t)

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
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
object, if it has one. The solids are assumed to be of uniform
density .

bn-1
Kn

K1 K2

Kn-1

4.6.3Mass Moment of Inertia Calculated from


Area Moment of Inertia

g
Fig. 4.45 Linear graph of the Maxwell viscoelastic model shown in
Fig.4.44

axis of symmetry is half that of the same mass at a distance


R from its axis of revolution, Fig.4.50.
J = L

R4
2

J =
L R 2
Cylinder
Volume

R2
2

J=M

R2
2

The parallel-axis theorem is used to calculate the mass moment of inertia of a solid which rotates about an axis parallel
to but at a distance from an axis for which the mass moment
of inertia of the solid is known. The parallel-axis theorem adds
the square of the distance between the mass center and axis of
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 = J Cylinder +
about its
axis

2
Mr


Parallel Axis
Theorem

=M

4.6.4Mass Moment of Inertia Calculated by


Superposition

Although the parallel-axis theorem is easy to remember, it


is not convenient to express the rotational inertia in terms of
mass, since any change in geometry also changes the mass. It
is more efficient, and generally provides better physical ap-

a
Ideal massless
and rigid bar

J xx = L I xx
(4.38)
where L is the length normal to the cross-section as defined
in Fig.4.55.

R2
+ Mr 2
2

4.6.2Mass Moment of Inertia of Primitive


Shapes

Fig. 4.46a Energetic schematic


of the Zener viscoelastic model.
b Zener viscoelastic model with
nodes of the across variable
translational velocity

The mathematical similarity between the area moment of


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
cross-section of a beam in resisting bending increases with
the square of its distance from the neutral axis, where the
axial stress is zero. This is mathematically analogous to the
kinetic energy in rotation increasing with the square of the
distance of a mass element from the axis of rotation.
If a machine component has uniform density and constant
cross-section, the mass moment of inertia J x x for rotation
about the x-axis normal to the cross-section through the centroid, and the area moment of inertia I x x for bending about
the z-axis through centroid along the length of the component are related as

The mass moment of inertia of solid volumes not shown in


the table can be calculated by superposition by adding and
subtracting mass moment of inertia of solid primitives or
with the parallel-axis theorem, after finding the mass and
centroid of the solid.

x,v

x,v
K1

K1

v(t)
K2

v(t)
g

K2

g
b
g

4.6 Modeling Rotational Elements

225

K2

K1

F(t)

g
Fig. 4.47 Linear graph of the Zener viscoelastic model shown in
Fig.4.46
Table 4.4 Analogies between translational and rotational mechanical
elements
Elemental equation Energy equation
Kinetic energy storage
Mass

FM = M

Rotational inertia

TJ = J

dv1g
dt

d 1g
dt

1
Mv12g
2
1
E J = J 12g
2

EM =

Strain energy storage


Translational spring

dFK
= Kv12
dt

EK =

FK2
2K

Rotational spring

dTK
= K 122
dt

EK =

TK2
2K

by removing two rings and the cylindrical shaft bore from


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 independent 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
would represent the motion using two rotational subsystems
with two different rotational inertias to store the kinetic energies due to the angular velocities about the two axes.
Pushing the web to the bottom surface of the hub and
flange allows the mass moment of inertia to be calculated as
either the sum of a positive cylinder plus a negative cylinder
and a negative ring.

4.6.4.1Example Mass Moment of Inertia


Calculation
Working with the cross-section of the pulley shown in
Fig.4.56 and the construction of the positive and negative
primitives shown in Fig.4.59a, the mass moment of inertia
equals the rotational inertia of the Positive Cylinder A minus
the rotational inertia Negative Cylinder B and twice the rotational inertia of Negative Ring A

J = J Cylinder A J Cylinder B + 2 J Ring A

Energy dissipation
Translational damper

Fb = bv12

None. Energy
dissipater

Rotational damper

Tb = b 12

None. Energy
dissipater

LA rA4 LB rB4
Lring 4

J =
ro ring ri 4ring

+ 2

2
2

J=
Example Calculate the rotational inertia of a pulley
shown in Fig.4.56 by the superimposition of the positive
and negative mass moment of inertia of solid primitives. The
pulley is 2024 aluminum with a unit weight, 0.101 lb/in 3 .
The dimensions are in inches. Report the result in SI units.
The mass moment of inertia can be calculated by summing
the rotational inertia of three rings, Fig.4.57.
This is not the only construction which yields the mass
moment of inertia. The same shape can be constructed
Fig. 4.48 Infinitesimal mass
element dm in a mass moment of
inertia at radius r from the axis
of rotation is accelerated by infinitesimal torque dT of applied
torque T

))

4
LA rA LB rB4 + 2 Lring ro4ring ri 4ring
2

Although it increases the time required for a computation, it


is best to convert units as a discrete step, because it permits
a reasonableness check before the results are obscured by
other operations, such as raising the dimension to the fourth
power, in this case. If the reader is thinking, I wouldnt
know whether the result of the unit conversion is reasonable or not, performing calculations of this sort is how one
develops a sense of scale

vi = r
H
T

dm
r

4 Mechanical Systems

226

We must reexamine the routine conversion from pounds to


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
units, force, length, and time, and SI, mass, length, and time.
Weight is the force exerted by gravity on a non-accelerating
mass

R
dr

r
+T,,

Fig. 4.49 Cross-section of a cylinder of length L and radius R normal


to axis of symmetry

rA =

6.5 in 2.54 cm
= 82.6 10 3 m
2 1 in

rB =

0.5 in 2.54 cm
= 6.4 10 3 m
2 1 in

r0 ring =

5.5 in 2.54 cm
= 69.9 10 3 m
2 1 in

ri ring =

1.5 in 2.54 cm
= 19.1 10 3 m
2 1 in

((

)(

25.4 10 3 m 6.4 10 3 m

+ 2 6.6 10 3 m

) ((69.9 10 m) (19.1 10 m) )
3

Converting only like to like units yields an apparently different result

kg m
1 N sec 2
1N
1kg =
=
2
m
m
sec
sec 2
Hence, our result is one gram shy of the conversion factor
0.454 kg/1 lb.
Unfortunately, US customary units are still used by the
aerospace industry. Unit conversion errors has led to costly
and well publicized errors, including the crash of a Martian
probe due the incorrect assumption that force was reported in
newtons when, in fact, it was reported in pounds. The most
unusual US customary units are possibly those used for mass
moment of inertia. The rotational inertia of DC servomo2
tors is sometimes given in units of [oz in sec ] leading to
question Is that ounce force or ounce mass? Do not waste
time searching for the appropriate conversion table. Perform
a unit analysis. We know that the units of mass moment of
2
2
inertia are [mass length ]. Attempt to convert [oz in sec ]
to SI units presuming that oz. represents ounce force,
1N = 1

lb 0.454 kg 39.36 in
kg
= 2800 3
3

in 1 lb 1 m
m

)(

W
g

Recall that 1 N is defined as the force needed to accelerate a


1kg mass at 1m/sec2,

25.4 m 3
3
Lring = 0.25 in
= 6.6 m
1 in

M =

N
N sec 2
= 0.453
m
m
sec 2

25.4 m
LA = LB = 1 in
= 25.4 m 3

1
in

kg

2800 3
m
25.4 10 3 m 82.6 103 m
J=

W = Mg

1lb 0.225 N 3.28 ft


= 0.453
32.2 ft 1lb 1 m

sec 2

= 0.101

F = Ma

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
L

J = L

r4
r2
=M
2
2

4.6 Modeling Rotational Elements

227

Fig. 4.51 Mass moment of


inertia of a right cylinder about
an axis parallel to its axis of
symmetry

r
R

R2
+ M r2
J = J Cylinder +
M r2 = M
2
about it
Parallel Axis

axis

Fig. 4.52 Mass moment of inertia (rotational inertia) J of solids


of uniform density

Theorem

Area A
Radius r

J=

A L3

r 2 L3

Length L

Right Cylinder Rotating About One End

Area A
Radius r

J=

2 A L3 2 r 2 L3
=
3
3

Length L

Right Cylinder Rotating About Its Center

Offset

Area A

J=

(
3

L + ) 3

Length L

Right Cylinder Offset from Axis of Rotation by Distance

ri
L

ro

J=

L
2

(r

4
o

ri 4 )

Right Cylindrical Tube or Ring Rotating About Its Axis of Symmetry

4 Mechanical Systems

228
Fig. 4.53 Mass moment of inertia (rotational inertia) J of solids
of uniform density

Radius r

x
z

J=

r 5
5

Hemisphere Rotating About Its Axis of Symmetry (y)


or an Axis Through the Center of its Base (x and z)
y

Radius r

r 3 2 r 4 r 5
J =
+
+
2
5
3

Offset
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

J=
x
z

r 5
5

Hemisphere Rotating About an Axis Tangent to its Pole


y

Radius r

r 5 2r 4 2r 3 2 r 2 ( r + )3
+

J = +

4
3
3
5

Offset
x
z

Hemisphere Rotating About an Axis Offset to a Tangent to its Pole

1m
2 1 1b 0.225 N
oz in sec 16 oz 1 lb 39.37 in

= 3.57 10 4 N m sec 2
We must introduce the definition of a newton in order to incorporate mass into the units,
kg m
1
3.57 10 4 N m sec 2 sec 2 = 3.57 10 4 kg m 2

1N

In this case, oz is ounce force. An engineer would use such


a strange unit, in order to avoid working with the inconveniently scaled magnitudes inherent to most SI units. Scientific notation requires remembering the multiplier and the
exponent. An error in remembering, or manipulating, the
exponent creates an order of magnitude error. US customary units are human scaled, meaning that the magnitudes
were greater than one, but not huge, making them easier to
compute and remember. The older metric system, cgs, for
centimeters-grams-seconds, also had this advantage. The results 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 inertia (rotational inertia) J of solids


of uniform density

J=

L
b

L
3

( a b + ab )
3

Rectangular Prism Rotating About an Edge

a 3b ab3
J = L
+

6
24

b
_a
2

_a
2

Rectangular Prism Rotating About the Center-Line of a Face

J=

L
3

( a (b + ) + a (b + ) )

a
Offset

L
3

( a b + ab )
3

Rectangular Prism Rotating About an Axis Offset from an Edge

a 3 ( b + ) a ( b + )3

+
J = L

24
6

a 3 a 3
L
+

6
24

b
_a
2

_a
2
Offset

Rectangular Prism Rotating About an Axis


Offset from the Center-Line of a Face
(millimeters) for that reason. Engineering notation has a
distinct advantage over scientific notation. By varying the
exponents only by a factor of three, values that are within
one or two orders of magnitude are likely to be expressed
using the same exponent. If a consistent exponent is used,
then to remember the number, one only need remember the
mantissa. Engineering notation also reduces errors made in
comparing the magnitudes of values, since one only needs to
compare the mantissas.

z
L
y
x
x

Fig. 4.55 The relationship between the mass moment of inertia J and
area moment of inertia I for a machine element with constant crosssection is J x x = L I x x

4 Mechanical Systems

230
Fig. 4.56 Plan and section of
pulley

0.25
1.0

0.5
1.5
5.5
6.5

Plan
Positive
Ring A

Positive
Ring B

Positive
Ring C

Section A-A

lindrical tubes. The rotational (or torsional) spring constant


is calculated using Eq.4.28 rearranged as


Fig. 4.57 Mass moment of inertia of pulley constructed of three rings

Rotational springs store elastic strain energy in torsion. The


common form of the equation for a torsion spring is analogous to the equation for a translational spring. The torque TK
acting through the spring is related to the angular displacement 12 across the spring.
TK = K 12

dTK
d
= K 12 = K 12
dt
dt

(4.40)

Angular deformation of an elastic material stores strain energy. Calculation of torsional deformation is quite complicated
except for the case of elements with circular cross-sections.
Fortunately in machine design, shafts and torsion bars are
usually circular in cross-section, either solid cylinders or cyFig. 4.58 Two geometries with
the same mass moment of inertia
about their axes of rotation

12

(4.41)

=G

(4.42)

where is the shear strain and is the shear stress. Shear


strain is an angular distortion of a material without volume
change. Strain is defined on the infinitesimal scale, Fig.4.61.
As the distance y approaches zero


(4.39)

Angular displacement is not a power variable in a rotational system, so, as with translational mechanical systems,
the common expression is differentiated with respect to time
to yield an expression in terms of the power variables torque
T and angular velocity

TK

The angular deformation 12 of an elastic material is calculated using the torsional analog of Youngs Modulus, the
Shear Modulus G defined as


4.6.5 Torsion Springs

K=

lim

y0

x dx
=

y dy

(4.43)

If a machine parts geometry is constant over a finite distance


y, and the shear modulus is constant in that volume, then we
can approximate shear strain on the macroscopic scale as


(4.44)

Shafts twist when torqued. A shaft with uniform shear modulus and constant cross-section will have a constant twist
angle, resulting in a straight line parallel to the axis twisting
into a helix when torque is applied. The twist of a shaft is not
the same as the rotation of a shaft. Twist is the deformation
of the shaft due to the torque transmitted through the shaft.
When a shaft carrying a constant torque rotates or spins to
transmit power, both ends rotate at the same velocity. It is

4.6 Modeling Rotational Elements

231

Fig. 4.59 Two equivalent mass


moment of inertia constructions
of positive and negative primitives

Positive
Cylinder A

Negative Positive
Ring B
Cylinder A

Negative
Rings A
Negative
Cylinder B

1 , 1

Helical line created by


constant twist angle

2 , 2

TK

TK

Node 1

Negative
Cylinder B

Angular displacement

Node 2
Straight line before twist

Fig. 4.60 Energetic schematic symbol for a torsion spring

b x

yx
y

xy

Fig. 4.62 A shaft with uniform circular cross-section and shear modulus will have a constant twist angle when carrying torque. Straight lines
parallel to the axis will be twisted into a helixes
y

xy

shear stress, (r)

yx
radius, r

Fig. 4.61 Strain deformation of an infinitesimal element under positive


shear stress . The subscript notation for shear stress 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.
The shear strain varies with distance from the axis of rotation on each cross-section. Shear strain equals the limit as the
distance between the cross-sections, L, approaches zero of
the angular displacement between two cross-sections of the
shaft, 12, divided by L and multiplied by the radius r from
the axis of rotation, Eq.4.34.


( r ) = lim

L0

r 12
L

(4.45)

Shear stress is proportional to shear strain, Figs.4.62 and


4.63. Hence, shear stress also varies with the radius,


( r ) = G ( r ) = G lim

L0

r 12
L

(4.46)

The torque applied to the shaft is balanced by shear stress


acting on cross-sections normal to the axis. Because the

Fig. 4.63 Distribution of shear stress on a plane normal to axis of


rotation and twist of a shaft

shear stress increases with the distance r from the axis of


rotation, the torque on an infinitesimal ring increases as r 2.
The infinitesimal torque at a distance r from the axis of
rotation is
dT = rdF = r ( r ) dA

where ( r ) = G ( r ) = G r , dF = dA, and dT = rdF ,


L
Fig.4.60.
The relationship between torque and angular displacement is derived by integrating the infinitesimal torque over
the cross-section, Fig.4.64. The surface area of an infinitesimal ring a distance r from the axis and of width dr is
dA =

r dr d = 2 r dr
0

4 Mechanical Systems

232

1 , 1

dr
r d
r

Tb

dT = rdF = rdA

Node 1

2 , 2

Tb
Node 2

Fig. 4.65 Schematic symbol for a torsional damper or drag cup

Rolling-contact bearing

b
Fig. 4.64 Distribution of shear stress on a plane normal to axis of
rotation and twist of a shaft

T(t)

T = dT = r dF = r dA = r r G dA

L
A
A
A
A
Mass-moment-of-inertia J
keyed or splined to the shaft

2 r 4

T = r r G 2 r dr = G

L
L 4
0
T =G

2 r 4
L

T = K

K=

Fig. 4.66 Schematic of a mass moment of inertia J keyed to a shaft,


which is supported on rolling-contact bearings with damping b and
driven by a torque source T(t)

G 2 r 4
L 4

This result contains yet another moment of inertia, in this case,


the polar moment of inertia, which describes how the distribution of area on a cross-section relative to the axis of rotation
contributes to creating an internal torque. Unfortunately, polar
moment of inertia has the symbol J. We will distinguish between polar moment of inertia and mass moment by identifying a polar moment of inertia with the subscript polar,


J polar

T=

2 r 4 r 4
=
4
2
G J polar
L

Rotational
Spring
Constant

(4.47)

(4.48)

4.6.6 Rotational Damping


Energy is dissipated in rotation by the same mechanisms where
energy is dissipated in translation: slip between surfaces, shear
of fluids, and deformation of viscoelastic materials. The same,
or a very similar, schematic symbol is used to represent damping in rotational systems. The rotational equivalent of a dashpot is a drag cup. Instead of fluid being forced to flow in
or out of an orifice or the annular space between a piston and
cylinder, as in a dashpot, fluid is sheared in the annular space
between a cylinder and an inner rotating drag.
Although energy dissipation is often deliberately designed into rotational systems, it is always present to some
degree, due to either fluid shear or deformation of viscoelastic materials. Unless a machine rotates in a vacuum, some

air will be pumped or moved by its motion. Shafts must be


supported on rolling-contact or hydrodynamic bearings or
on simple bushings. These devices have both viscous and
Coulomb friction.

4.6.6.1 Rolling-Contact Bearings


Another common schematic representation of a rotating
element supported on rolling-contact bearings is shown in
Fig.4.66. The drawing shows ball bearings.
Other types of rolling-contact bearings include roller, tapered roller, and needle bearings.
The shaft is supported on two rolling-contact bearings.
The rolling-contact bearings are shown in cross-section with
the plane of the section through the axis of rotation. Often,
the cross-sections sections of rolling-contact bearings are not
drawn and simply indicated by Xs. The rotating element is
indicated to have rotational inertia J. The shaft and rotational
inertia are typically two separate elements keyed, splined, or
coupled together. Each bearing is indicated to have damping
coefficient b. The rolling friction of steel on steel due to the
deformation of the shaft and rolling-contact bearing is negligible, but the seals contribute Coulomb or dry friction of
some amount. Bearing are lubricated with either oil or grease
depending on their use. Each bearing has damping b due to
the shear of the lubricant and damping due to the motion of
the rotating element it supports in air or a more viscous fluid.
The fits and preloads placed on bearings vary, but, in general, the inner race of a rolling-contact bearing is press fit or
a shrink fit onto the shaft and rotates with the shaft. The outer
race of a rolling-contact bearing is press fit into the bear-

4.6 Modeling Rotational Elements

233

a
b
T(t)

T(t)

J
1

Hydrodynamic
Bearing

g
The outer bearing race
is fixed in the bearing mount

Fig. 4.67 Nodes of the across variable angular velocity shown on the
energetic schematic. The inner race of the rolling-contact bearing rotates at the shafts velocity. The outer race is fixed in the bearing mount
and does not rotate

Ra

Shaft

diu

sr

v1 = 1r
vg
Bearing

Stationary Bearing g

T(t)

J
b

1
b

The inner bearing race


rotates at the shaft speed

Rotating
Shaft

Fluid Velocity
Profile

Fig. 4.68 Cross-section through a hydrodynamic bearing in a pillow


block. The ideal fluid velocity profile is linear between the two solid
surfaces. A typical gap between the shaft and bearing is 0.0010.003in.

ing mount or pillow block and does not move relative to the
mount, which is often in the machines frame and at ground
velocity. In the annotated figure the outer races of the bearings are assumed to be fixed to the machine frame, which is
an inertial ground.
Again, note that the inner race of bearings rotates with the
shaft the bearing supports, and the outer race is stationary
relative to the machine frame. Consequently, the angular velocity drops across the bearing from the velocity of the shaft
to ground velocity.

4.6.6.2 Hydrodynamic Bearings


Hydrodynamic bearings are used to support large forces. The
rotation of shaft shears and drags the lubricating fluid. The
drag pressurizes the lubricant, allowing the fluid to support
the transverse force of the shaft and prevent metal to metal
contact. Their most familiar application is as the crank shaft
and piston rod bearings in automobile engines. A hydrodynamic bearing cross-section is shown in Fig.4.68. The bearings are made of a steel, copper or, aluminum ring overlain

Fig. 4.69 Abstract schematic of a torque source-rotational inertiadamper system, where the damping is represented as a hydrodynamic
bearing supporting the rotational inertia

by soft tin alloy known as babbit, after its nineteenth century


inventor, Iaasa Babbitt. The soft alloy protects the surface
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
by rolling-contact bearings. Although rolling-contact bearings have a Coulomb friction torque of some magnitude due
to seals and packing, we will generally model bearings as
ideal hydrodynamic bearings with a linear relationship between the angular velocity of the shaft relative to the bearing
and the torque needed to maintain that velocity.
In most cases, the damping in a rotational system is determined experimentally. In those cases, all of the damping
in the system is reported using a single damping coefficient
and, if necessary, a single static Coulomb friction torque.
An abstract energetic schematic representation of a rotational system with a bearing is shown in Fig.4.69. This representation makes no attempt to represent the configuration
of either the rotational inertia or the type and number of bearings supporting the shaft. It simply indicates the presence of
those energetic elements in the system by their parameters,
b and J. Note that the torque source acting against the rotational inertia must also react against ground, Fig.4.69b.
Schematics of rotational systems in which two or more
energetic attributes have the same angular velocity drop
across them and act in parallel can be difficult to draw and to
interpret. If the attributes are possessed by a single machine
element, or when the attributes are possessed by machine
elements which are coaxial, then, in a two dimensional drawing, the components lie atop and obscure one another, unless
they are drawn in a cross-section normal to the axis of rotation. When schematic symbols of two attributes, say a torsional spring and a drag cup or rotational damper, are drawn
parallel to one another, so that each is visible, then it appears
that they could not rotate at the same angular velocity.

4.6.6.3 Fluid Coupling


A fluid coupling is a device which transmits torque by shearing fluid captured between two solid components rotating at

4 Mechanical Systems

234

Tb
Node 1

400

2 , 2

Tb
Node 2

Fig. 4.70 A fluid coupling symbol. The drag cup symbol is more
common

different angular velocities. The most familiar fluid coupling


is the torque converter of automatic automobile transmissions, in which vans on one half of the housing pump automatic transmission fluid (oil) at vans on the other half of the
housing. The prime advantage of a fluid coupling is that it
allows one of the shafts to have zero velocity, while the other
shaft continues to rotate. In its automotive application, this
allows the engine to run while the drive shaft and the car are
stopped. The disadvantage of a fluid coupling is the energy
lost in shear of the working fluid. The automatic transmission fluid must be cooled to limit its temperature rise. Most
automobiles have a cooling coil inside the radiator, which
connects to tubes from the transmission. Older automatic
transmissions dissipated 10% of the power input as heat.
Modern automatic transmissions can be locked-up at highway speed with the engine power transmitted through gears
to increase fuel efficiency.
An actual torque converter is roughly toroidal in shape.
The schematic symbols used for a fluid coupling is either a
drag cup, Fig.4.65, or two L-shaped lines, Fig.4.70, where
the gap between them is the fluid-filled volume, which is
sheared. The drag cup symbol is more commonly used.

4.6.7 Rotational System Sources


Shafts are the most common rotational system power sources. During the industrial revolution, entire factories were
powered by a single steam engine and the power distributed
throughout the building by overhead shafts. The legacy of
this is seen today on the campuses of older colleges and universities, where part of the mechanical engineering department remains near the steam plant.
Rotational sources are machines or machine elements in
which we can control or specify one of the two rotational
power variables, either torque or angular velocity. An example of a rotational source is a rotating shaft separated from
the rest of the rotational system by a friction clutch. If the
machine driving the shaft is powerful enough to approximately maintain the shafts speed when the clutch plates are
closed, then we could model the machine-shaft-clutch as an
angular velocity source. Note that this criterion depends on
the amount of power the source can supply and relative to
that the system draws and can only be satisfied approximately. Remember, we cannot control the power drawn from the
source. Consequently, we can only control one of the two

Percent of Rated Torque

1 , 1

Design A

Design D

300
200

Design C

100
0

Design B
0

20

40

60

80

Percent of Rated Speed

100

Fig. 4.71 NEMA three phase induction motor standard designs A, B,


C, and D torque vs. speed curves

power variables. The magnitude of the second power variable is determined by the amount of power the system draws
from the source at any instant.
Torque sources are generally transducers of some sort, in
which we control the power variable, which transduces into
torque. Using a positive displacement hydraulic motor as an
example, fluid flow rate is proportional to angular velocity
of the motor. Consequently, fluid pressure is proportional to
torque. If we can (approximately) control the fluid pressure
entering the hydraulic motor, then we can control the torque
output. Likewise, in an electric motor, torque is proportional to motor current, and angular velocity is proportional to
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
device which modulates (or changes) the voltage applied
to the system to maintain a specified current. Current sources are available and used in industry specifically so that the
torque an electric motor produces can be controlled.
AC induction motors are widely used in industry. They
have no brushes, which are actually graphite contacts, and
require less periodic replacement than other types of motors.
There are four standard designs of three phase induction motors are available, NEMA (National Electrical Manufacturers
Association) designs, A, B, C, and D. The designs provide
different torque-speed curves, as shown in Fig.4.71. The
torque curve of design C is relatively flat over most of its
speed range and approximates a torque source.
A design advantage of induction motors is their torque
at zero angular velocity. Many rotational motors must rotate
to function. Internal combustion engines are an example. If
the crank shaft stops rotating, the engine stalls and must
be restarted using the auxiliary, electrically powered starter
motor. Likewise, a synchronous AC motor cannot produce
torque, except when the rotor is in synchronous rotation with
the rotating magnetic field of the stator. If a synchronous AC
motor falls out of synchrony, an auxiliary motor must bring

4.7 Dynamic Tests

235

max
Torque-Velocity Data

T0

Torque

Constant
Torque
Model

Constant
Velocity
Model

Tmax
0

Angular Velocity

Fig. 4.72 Constant torque and constant velocity models and their operating ranges

back up to speed. Stepper or stepping motors can produce


a holding torque at a constant angular position. Some rotational hydraulic motors can also produce torque without
rotation. We will consider motors in Chaps.6 and 11.
The operating ranges of the sources torque and angular
velocity must both be considered, when modeling a rotational power source. If relatively little power is drawn from the
source by the system, then it may be reasonable to model a
source as either a constant torque or a constant angular velocity source, Fig.4.72. If the amount of power drawn from the
source affects the magnitude of the variable, which we would
like to model as under our control, then we must add a damper in parallel with the source to represent its behavior. The
damping coefficient b is the ratio of the maxima of the torque
and the angular velocity, where they are chosen to provide
the best fit through the sources performance data, Fig.4.73.

4.7 Dynamic Tests


Dynamic tests are performed to establish the time-dependent parameters of viscous fluids, viscoelastic solids;
to characterize components of energetic systems; and to
confirm or validate an energetic model. The parameters
of the individual elements of energetic models are often
established through testing of a component or the entire
actual physical system, when it is feasible. Viscous damping and Coulomb friction are present to some extent in all
Fig. 4.73a Torqueangular
velocity data for a torque source
with two linear approximations
or models. bTorque source with
a damper in parallel to yield a
torqueangular velocity model.
This is a Norton source model. A
Thevenin model is a damper in
series with a source

mechanical systems. These phenomena are very difficult to


predict from fundamental principles due to the large effect
of small geometric features and other factors, as discussed
in Sect.4.2.4.
Viscoelastic materials are not characterized by convention tensile tests, because their properties are time
dependent. Three tests are commonly used. A creep test
measures the displacement of the material under a fixed
load. A relaxation test measures the decay of stress under
a fixed displacement. Dynamic mechanical analysis subjects specimens to sinusoidal loads to measure their frequency response, the subject of Chap.10.
Ultimately, tests must be performed to validate a model.
A models value is its ability to predict the behavior of the
physically real system. Until a models predictions are tested,
they remain open to question. Some models cannot be tested
by subjecting an exemplar of the system to a full scale test.
The effect of earthquake ground motions on a large structure is an example. In those cases, subsystems or small scale
physical models of the entire system are tested.
Dynamic test data are interpreted by equating an observed
value with a corresponding term from the system equation,
or by fitting a response function to the observed response
to test inputs representative of the systems operating range.
Parameters for linear models of energetic systems are estimated by measuring characteristic times, i.e., time constants
or periods of oscillation, and initial and final values of observed responses. Dynamic testing of viscoelastic materials
is more involved, due to the number of superposed Newton or KelvinVoight elements of different relaxation times
(or time constants) needed to approximate the response.

4.7.1Components with a Single Unknown


Energetic Parameter
If an individual machine component has a single significant
energetic property, then that parameter can be estimated
from the components steady-state response to a load (force
or torque) or velocity input, if it is feasible to remove the
component from the system, and the means exist to test it.

b
Tmax

Torque-Velocity
Data

Torque

Tmax
b = ___

T(t)

max

Torque-Velocity
Models

Angular Velocity

max

System

4 Mechanical Systems

236
Fig. 4.74 Load and displacement histories of a linearly
elastic component

Extension
xa = x1- xg

Force
Steady-State
ta

The most familiar example is loading an elastic machine


component to estimate its translational spring constant K,
Fig. 4.74. If a material is reasonably modeled as linearly
elastic, then the deformation of the material is proportional
to the force acting through it, where the proportionality constant is the spring constant K,

FElastic FK = K x1 xg = Kx1g K =

x1g
FElastic

Although less familiar, the torsional spring constant K of


a shaft or rod is measured similarly, by applying a known
torque and measuring the resulting angular displacement.
Although we commonly speak of applying a load, either
a force or a torque, to an elastic element or structure, we cannot instantaneously apply a fixed load. It would require infinite power. Specifically, the testing machine would need to
displace a finite distance at infinite speed. Consequently, we
must build up the load since an elastic element deforms as
it is loaded. Typically, a load is ramped up at a uniform rate,
Fig.4.70. The test should span the expected operating range
of the component. In theory, we could use any pair of displacement and load data to determine the spring constant K. In
practice, the spring constant is calculated using the expected
force which the component will most often experience, that is,
its operating point. The initial portion of the load history, near
the base of the ramp, is not used, because there is slack in the
load path, which is pulled or twisted out at the start of a test.
The damping coefficient b of a translational or rotational
damper is also determined by load tests which span the expected operating range. Depending on the magnitudes of the
loads and velocities involved and the capability of the testing
equipment, the tests may be either continuously varying velocities or repeated step inputs of different velocities. In either
case, the velocity and the corresponding force or torque should
span the expected operating range of the damper. Amateur
auto racing enthusiasts characterize the frequency response
of shock absorbers using an inexpensive shock dyno, which
is a crank and slider mechanism driven by a variable speed
electric motor. Frequency response is the subject of Chap.10.
If the Coulomb and viscous friction acting on a mass or a
mass moment of inertia are negligibly small, then the mass
M or mass moment of inertia J are determined by measuring
their acceleration under a known force or torque. It is most
common, however, for the Coulomb and viscous friction to

Time

Steady-State
ta

Time

be non-negligible. In this case, the mass or mass moment of


inertia and its associated friction are estimated by using both
the transient and steady-state portions of the test response, as
discussed in the following section.

4.7.2Components with Multiple Energetic


Parameters
The discussion above presumed that a machine component
had a single energetic property, and the property could be
calculated from material test data, or by a test of the machine
component itself. The most common circumstances are that
(1) it is either not feasible or downright impossible to remove a machine component from a system for testing, or (2)
a single machine component has two or more significant energetic properties. In these cases, the unknown parameter or
parameters must be determined from the dynamic response
of the system to test inputs. Three different test inputs are
used: an impulse, a step, and a sinusoid. We will defer discussion of the response of systems to sinusoidal inputs, until
we discuss frequency response, but, briefly, the steady-state
magnitude and phase shift of the systems sinusoidal response provide two equations and, consequently, generally
allows us to solve for two unknown parameters, unless we
have more unknowns or the unknowns appear in ratios.
The interpretation of impulse and step response data derives directly from the response equations expressed in terms
of the input and the system parameters. In theory, an impulse
input produces the homogeneous response of a system. The
steady-state, or particular, response of a system to an impulse
is zero since the impulse is zero, except at the instant it turns
on. There are two practical difficulties. First, the ideal unit
impulse, (t ), is a singularity function defined to have zero
duration (or width on the time axis), but its time integral (or
area) equals unity, which requires its magnitude to be infinity at the point where its argument equals zero. There are
no macroscopic physical phenomena which possess these
properties that we can use as inputs. The most useful tool for
approximating an impulse input is a large hammer which is
used to excite vibrations in structures. A swinging pendulum
fills the role of a large hammer for impact testing.
In practice, a step input is often the most convenient and
realistic test input. Step inputs are convenient, because many
of the systems we design are on or off, that is, either

4.7 Dynamic Tests


Fig. 4.75a Force source acting
on a spring K and damper b.
b Linear graph of the system
shown schematically in a

237

Rigid and
massless bar

F(t)
x,v

F(t)

4.7.3 First-Order System Step Responses


The transient portion of a first-order systems response provides the systems time constant which is a ratio of the damping coefficient and the energy storage parameter. The initial
or steady-state value may be a product or ratio of the input
and the damping coefficient.
Polymers (including most biomaterials), amorphous
(or glassy) materials, and crystalline metals at sufficiently
high temperatures exhibit time-dependent deformation at
constant stress. This phenomenon is known as creep. The
combination of stress-dependent and time-dependent deformation is called viscoelasticity. The size of polymer molecules leads to significantly different mechanical behavior
than that of metals and ceramics. Viscoelasticity in polymers
is due to the slippage of the macromolecular chains past
each other. The slippage is impeded by side groups attached
to the main molecular backbone. Creep of viscoelastic materials increases with temperature, because the thermal motion of
the molecules increases the likelihood that side groups which
impede slippage will vibrate out the way of one another.
Manufacturing a dog-bone specimen from the same
material introduces the possibility that the effects of the
manufacturing processes used to make the actual parts and
those used to make the specimens have different effects on
the materials behavior. Solidification processes such as casting or molding; deformation processes including stamping
and forcing; and material removal processes such as turning or milling, all have stress, deformation, and temperature
histories which depended on the size and shape of the part.

g
connected or disconnected to a source with a constant power
variable. We can perform a test by operating the system in
its normal mode with its own power supply. For on-off
systems, this type of test is clearly also the most realistic,
and, consequently, most likely to lead to a characterization of
the systems parameters which are useful in dynamic modeling. A step input has the additional advantage of providing
non-zero steady-state data, in many cases.

The material properties of a tensile specimen can never be


exactly those of the corresponding machine component.
The question is always, How close are the properties of the
test specimen to that of the part?
To a degree, all material testing are index tests, where
the results of the tests correlate with the performance of
parts, rather than provide material properties. The underlying 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
the data, all contribute to making material characterization
challenging. In order to use any of the test results within a
linear model of an energetic system, we must select a value
of the parameter to represent a population of data, limiting
the precision of the model.

4.7.3.1Estimating Spring Constant K and


Damping Coefficient b from Step
Response Data
We will use the first-order force source-spring-damper system shown in Fig.4.75 as the first example. The energetic
equations, Sect.3.1.4.1, and the system equation for the force
FK acting through the spring, Eq.3.29, are reproduced below.
Continuity or Node Eq:

F (t ) Fb FK = 0

Compatibility or Path Eq:


Element Eqs:
Energy Eqs:

Fb = bv1g

E sys = E K
F (t ) =

v1g = v1g
dFK
= Kv1g
dt

EK =

FK2
2K

b dFK
+ FK
K dt

(3.29)


The step response data are the displacement, x1g, Fig.4.76.
We easily derive the system equation for the displacement,

4 Mechanical Systems

238

12

12

10

10

x1g(t)
cm 6

x1g(t) 6
cm
4

0.5

1.0

1.5

t, sec

2.0

Fig. 4.76 Displacement in centimeters under a step input of 4,000N

x1g, by eliminating the spring force FK from Eq.3.29 with the


substitution FK = Kx1g , Eq.1.6,
F (t ) =

x1g()=7.2 cm

2
0

x1g (ss)=11.4 cm

b dFK
+ FK
K dt

F (t ) =

b dKx1g
+ Kx1g
K dt

1
b dx1g
+ x1g
F (t ) =
K
K dt

(4.49)

t
K
t

x1g (t ) = x1g ( ss ) 1 e x1g (t ) = 0.114 m 1 e b

The input F (t ) = 4, 000 N us (t ) is constant in steady-state.


Consequently, the derivative of the output variable is zero in
steady-state, allowing us to determine the spring constant K

1.5

2.0

Fig. 4.77 Steady-state displacement equals 11.4cm. The displacement


after one time constant equals 63.2% of steady-state, 7.2cm. The corresponding time is the time constant 0.26 sec

shown in Fig.4.77. The corresponding time t equals one time


constant . For these data, 63.2% of dynamic range is
0.632 11. 4 cm = 7.2 cm

The step response is a stable first-order growth

1
b dx1g ( ss )
F ( ss ) =
+ x1g ( ss )
K
K
dt
0

0.5
1.0
0.26 sec t, sec

1
F ( ss ) = x1g ( ss )
K

which occurred at approximately 0.26sec. The time constant


is found by unit analysis of the system equation
1
b dx1g
F (t ) =
+ x1g
K
K dt


b K 0.26 sec

b 35, 000

b 9,100

b
0.26 sec
K
N
0.26 sec
m

N sec
m

Alternatively, we can work from the step response function

F ( ss )
4, 000 N
N
=K
= 35, 000 = K
x1g ( ss )
0.114 m
m

t
K
t

x1g (t ) = 0.114 m 1 e x1g (t ) = 0.114 m 1 e b

To determine the damping coefficient b, we must first


estimate the systems time constant. We can calculate the
time constant using any data pair, the initial value and the
dynamic range. However, the calculation is more accurate, if
we use data from early in the response, where rate of change
is greatest. A first-order step response progresses 63.2% of
the remaining change toward its steady-state value in one
time constant .

Only positive numbers have logarithms, so, dropping the


units of meters, we rearrange the response function as

(t ) = ( ss ) 1 e

(1)
(1 ) = ( ss ) 1 e

( ) = ( ss )(1 0.368) ( ) = ( ss )( 0.632)


A convenient method is to calculate 63.2% of dynamic range
and then locate that value approximately in the data set, as

K
t
b

0.114 x1g (t )
0.114

where the value of 0.114 x1g (t ) is greater than zero since


0.114 = x1g ( ss ) . We can now take the natural logarithm
0.114 x1g (t )
Kt
ln e b = ln

0.114

0.114 x1g (t )
K
t = ln

b
0.114

4.7 Dynamic Tests


Fig. 4.78a Cross-section of a
flywheel with mass moment of
inertia J supported on rollingcontact bearings driven by a
torque source T(t). b Linear
graph of the rotational system
modeling the energy loss from
the bearing and air drag as ideal
viscous friction with damping b

239

Mass moment of
inertia J

Rolling-contact
bearing

T(t)

T(t)

g
g
2,000

Since the spring constant K is known, the expression above


can be evaluated using any data pair, other than one in
steady-state, to calculate the damping coefficient b.

4.7.3.2Estimating Torque and Damping from


Step Response Data
The objective is to determine the energetic parameters of
a rotational system consisting of an induction motor driving a flywheel on a shaft supported by rolling-contact bearings, Fig.4.78. The mass moment of inertia of the flywheel
and the damping of the system are unknown. The induction
motor driving the system has one of the torque vs. speed
curves shown in Fig. 4.71, but the design type of the motor
is unknown. A test is performed in which the motor is energized and the angular velocity of the flywheel is measured,
Fig. 4.79. Note the response curve resembles a linear firstorder step response, but it does not look quite right. Either the
damping of the system is not linear or the input is not a step.
In this example, the damping is linear. The torque applied by
the motor is not constant. We wish to make an initial estimate of the damping coefficient and the input torque. We will
model the input torque as a step input of unknown magnitude,
i.e., a constant torque, even though we know that it is not.
Fig. 4.80a Cross-section of
the flywheel. b Cross-section
of equivalent geometry used for
calculation of the flywheels volume and mass moment of inertia

1,500

(t)
1,000
RPM
500
0

20

40

60

t, sec

80

100

120

Fig. 4.79 Step response of the system to an unknown torque input

The dimensions of the flywheel are shown in Fig.4.80.


The flywheel is either iron or carbon steel. Fortunately, the
weight of the flywheel is known to be 360 lbs. The shaft is a
carbon steel, 1.5in. in diameter and 17in. long.
The dimensions and weight of the flywheel are sufficient to
calculate the flywheels mass moment of inertia. The metals
density is needed for the mass moment of inertia calculation. The flywheels volume can be computed by using either
rings or solid cylinders. The flywheels geometry, Fig.4.80a,

b
2.5 in

1.5 in

3.5 in

16 in

2.5 in

1.5 in

3.5 in

2.5 in

2.5 in

2 in
4 in
10 in

16 in

2 in
4 in
10 in

4 Mechanical Systems

240

can be constructed by summing three rings; the flange, web,


and hub. Figure4.80b was drawn from Fig.4.80a by shifting
material parallel to the axis of revolution, leaving the volume
and mass moment of inertia unchanged. Figure4.80b can
be constructed by summing two positive and two negative
cylindrical volumes; the outer cylinder minus the inner cup
plus the hub projection minus the shaft bore.
The volume is 1, 260 in3 or 0.0207 m3. The density of the
metal is
360 lb
lb
= 0.285 3
1, 260 in 3
in
360 lb 0.454 kg 1 in
kg
= 7,890 3
3

1 lb 0.0254 m
1, 260 in
m

The material is likely a carbon steel.


The mass moment of inertia will be calculated by superposing the mass moment of inertia of cylinders rotating
about their axis, Fig.4.50,
J cylinder = L

J=

d
T t
T
[ J ] = = [T ] = [b ] [b ] =

dt

[T ] = J

Check the units of the system equation


T (t ) J d 1g
b = b dt + 1g

The units check.


We need the steady-state value and the time constant,
both of which we determine from the system equation in
time constant form
T0 J d 1g
=
b b dt

1g ( ss ) = 1, 670

E sys = E J

In reality, the motors torque is a function of its angular


velocity. The motor will reach a torque equilibrium with the
system in steady-state,

Tb = b 1g

T ( ss ) = b 1g ( ss )

1
E J = J 12g
2

However, in order to make any estimate, we must model the


torque as constant.
The time constant of the system is

Reduction: Input T(t), Output 1g(t)

T (t ) = TJ + Tb

T (t ) = J

d 1g

T (t ) J d 1g
=
+ 1g
b
b dt

rev 2 rad 1 min


rad
= 175
min rev 60 sec
sec
T0
rad
= 175
b
sec

1g = 1g

dt

The steady-state velocity is known, 1g ( ss ) = 1, 670 RPM.


Rotational speed in RPM must be converted to angular velocity in radians per second calculations in SI units

Continuity ( Node ) Eq: T (t ) = TJ + Tb

d 1g

T0
= 1g ( ss )
b

Energetic Equations

Energy Eqs:

+ 1g ( ss )

4
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 )

2
in

Compatibility ( Path ) Eq:

[ ] = [ ] + [ ]

r
2

To proceed, we need the system equation relating the input


torque to the angular velocity of the flywheel.

TJ = J

T J
b = b t + [ ]

J = 4 . 41kg m 2

Element Eqs:

T t
T T = T t + [ ]

Express the parameters J and b in terms of the power variables and time

dt

+ b1g

T0 J d 1g
J
=
+ 1g =
b
b dt
b

We estimate the time constant from the data as if the response were a linear step response, by locating the time in

4.7 Dynamic Tests

241

250

1,600
1,200

(t)
RPM 800

150

1g (ss)=1,670

(t)
___
rad 100
sec

1g ()=1,060

400

Motor torque

T(t) 200
N m

Flywheel velocity

50

t, sec

3.18 sec

10

15

Fig. 4.81 Time constant estimate

t, sec

10

15

Fig. 4.83 Response of the linear model to the variable input torque of
an AC induction motor

1,600
1,200

(t)
RPM 800

Linear model
Observed response

400
0

t, sec

10

15

Fig. 4.82 Linear model plotted against the observed response

the response at which the system has reached 63.2% of its


steady-state velocity, Fig.4.81,
0.632 1g ( ss ) = 1g ( ) 0.632 1, 670 RPM = 1, 060 RPM
We can now estimate the damping coefficient b,
J
b

=
b=

4.41 kg m 2
3.18 sec

b=

b = 1.39

N m sec
rad

the torque input T0,


T0
rad
= 175
b
sec
N m sec
rad

T0 = 1.39
175
= 243 N m

rad
sec
We know that these estimates are in error to some degree.
Plotting the ideal, linear step response calculated using these
estimates will provide a sense of how large the error is,
Fig.4.82. The step response function is

1g (t ) = 1g ( ss ) 1 e

1g (t ) =

T0
b

b
t

J
1
e

The observed response was created using a MATLAB program of the RungeKutta finite difference solver, using a
look-up table for the inductor motors torque, discussed in
Chap.8. The value used for the damping coefficient b was
b = 0.75 Nmsec/rad . The calculation overestimated the
damping coefficient by 85%. Correspondingly, the estimated torque applied by the motor, T0 = 243 N m, is also high.
The maximum torque applied by the motor was 239 N m,
Fig.4.83.
The errors in the estimated damping coefficient and torque
are due to the induction motors torque varying with its angular velocity. The initial portion of the response is driven
harder than later portion, decreasing the estimated time constant and leading to the overestimated damping ratio.
The initial response of many students is to view the
magnitude of the errors in the estimated values as outrageous and to see no value in the calculation. However, it
is important to note that prior to the calculation, we had no
inkling of the magnitude of either the torque or the damping coefficient. We estimated a value of the motor torque,
which is just outside the range of the actual torque. The
input variables and parameter values presented in textbook
problems must be determined by the engineer in practice.
Torque measurements are a particular problem. A torque
cell of the correct capacity must be inserted into the power
train. This is straightforward in a laboratory situation but
not when there is an existing system which cannot be taken
out of service and torn apart. Conversely, angular velocity is simple and cheap to measure. A hand-held laser tachometer is less expensive than most engineering students
calculators.

4.7.3.3Estimating Mass Moment of Inertia,


Damping, and Coulomb Friction from
Step Response Data
Energy dissipation in mechanical systems is due to shear
displacement between surfaces or within a plastically deforming material. The linear model of ideal viscous friction is used to approximate velocity-dependent shear force
or torque. Often, a shear force or torque which is not ve-

4 Mechanical Systems

242

Winding in a helical slot. Rotor


constructed of a stack of stamped
plates of electrical (silicon) steel

Magnetic flux emerging


from the rotors north pole

-i Graphite brush

Commutator

+i

Energizing
current i

Fig. 4.84 Schematic of the rotor, commutator, and brushes of a DC


motor

locity dependent must be included for the energetic model


to reasonably approximate the systems behavior. One
mechanism is Coulomb friction due to surface-to-surface
contact. Other phenomena which contribute a constant
shear or torque force are Bingham fluids, which have shear
strength at zero shear velocity, and electromagnetic forces
which may act on mechanical elements in electrical machinery, i.e., magnetic cogging in motors due to residual
magnetic fields.
Regardless of the physical mechanism, which creates the
shear force or torque required to displace two solid surfaces, it is modeled as if it were Coulomb friction, Fig.4.17. A
source must be used to exert a constant force or torque on
a system. The need to use a source to apply a shear force
or torque limits the operating range to one direction of motion, since friction acts to oppose motion. If there is an oscillation or vibration, and the motion of a mass or rotational
inertia reverses direction, then the friction force must also
reverse direction. This non-linearity of behavior is impossible to include in a linear model. We shall see, however,
that it is easy to include such action, when we work with
numerical methods in Chap.8. For the meantime, we must
restrict models with Coulomb friction to motion in one direction. Further, we must not accept results in which the
Fig. 4.85a Schematic crosssection of the DC motors rotor
with mass moment of inertia
JM. The shaft is supported on
rolling-contact bearings. The
bearings and air resistance of the
irregular shape of the motors
rotor are combined as damping
bM. Kinetic Coulomb friction TC
acts on the shaft. b Linear graph
of the mechanical aspect of the
DC motor

friction force or torque applied by the source drives the


system. This nonsensical behavior is predicted, when an
initially energized system is decelerated under a constant
friction force or torque. The models 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.
A DC servomotor is a motor used in robotics and industrial
automation. We shall investigate electric motors in Chaps.6
and 11. Briefly, servo derives from the Latin for slave
and refers to a device or a system under feedback control. A
servomotor has a tachometer and, often, a position encoder
built in to provide feedback from the motor to the controller. DC motors and series wound AC motors, which are used
in power tools, have graphite brushes, which are pressed
against a rotating contact, the commutator, to progressively
energize the windings on the motors shaft, Fig.4.84. The
motor rotates to align the magnetic field of rotor with the
stationary magnetic field on the motors frame or stator. As
the rotor approaches the equilibrium position, contact is broken with the energized winding and the adjacent winding is
energized, repeating the process.
We will derive a linear model using a linear graph, in
which the kinetic Coulomb friction TC acting on the motors
shaft is represented as a constant torque applied by a source,
Fig.4.85. We wish to estimate the mass moment of inertia of
the motors rotor, JM, the linear viscous friction (damping)
coefficient, bM, and kinetic Coulomb friction acting on the
motors shaft, TC. If the motor is energized and brought up
to speed, storing kinetic energy in the rotors mass moment
of inertia, and the electric circuit opened (the power turned
off), then only the mechanical attributes of the motor can
contribute to the response, since the electrical subsystems is
de-energized. The motor coasts down and stops. If the motor
were comprised of ideal linear mechanical elements, the
change in angular velocity, as it coasts down from an initial
angular velocity, would be an exponential decay. However,
the Coulomb friction acting the shaft changes the shape of

a
Rolling-contact bearing
linear damping, bM

TC

Kinetic Coulomb
friction torque

Motors rotor mass


moment of inertia JM

TC

JM

g
g

bM

4.7 Dynamic Tests

243

300

1g (0)

(t) 200
rad
___
sec

0.632 1g

1g (t)

1g

1g ()

100

1g (ss)

0
-0.2

0.2

0.4

0.6

t, sec

0.8

t, time constants

Fig. 4.86 Angular velocity of a DC servomotor after the electric circuit is opened as the motor coasts to a stop

Fig. 4.87 The linear model of the coast-down of a DC motor due to


kinetic Coulomb friction and a linear viscous friction predicts a nonsensical negative steady-state angular velocity

the observed response, Fig.4.86. The abrupt stop is due to


Coulomb friction.
Energetic Equations

The response function is a decay from a positive initial value


to a negative steady-state value

Node:

TC = TJ M + TbM

Loop:

1g = 1g

Elements:
Energy:

TbM = bM 1g

E system = E J

TJ M = J M

EJ

d 1g

dt
1
= J M 12g
2

TC = J M

d 1g

+ bM 1g

dt

TC
J d 1g
= M
+ 1g
bM
bM dt
The motors time constant is
TC
J d 1g
= M
+ 1g
bM
bM dt


JM
bM

Although the Coulomb friction torque TC is the input to this


model, the value of TC is unknown. We know the signs of the
unknowns. Parameter values, JM and bM, are always positive.
Coulomb friction acts to oppose motion. If we define the initial angular velocity of the system, 1g (0), to be positive,
then TC is negative. The steady-state velocity predicted by
the model is
TC J M d 1g ( ss )
+ 1g ( ss )
=
bM bM
dt
0

1g (t ) = 1g ( 0) 1g ( ss ) e

Reduction: Input: TC, Output: 1g


TC = TJ M + TbM

1g (t ) = 1g e + 1g ( ss )

TC
= 1g ( ss )
bM

bM
t
JM

+ 1g ( ss )

The models response, plotted in Fig.4.87, predicts that the


rotational inertia of the DC motor will coast-down to a fictitious negative steady-state velocity 1g(ss) due to the constant Coulomb friction torque TC. Consequently, the results
are only valid for 1g>0. We will not know the fictitious
steady-state value 1g(ss). Consequently, we cannot calculate
the dynamic range of the model 1g = 1g (0) 1g ( ss ).
Without the dynamic range, we cannot estimate the time
constant . At best, knowing that 1g > 1g (0), we can
establish a lower limit for the time constant.
We have three unknowns, TC, JM, and bM,which appear in
ratios in the response function.

T
JM
and C = 1g ( ss )
bM
bM

Having two equations and three unknowns, there is not


enough information from the coast-down test of the motor.
Changing the initial angular velocity and repeating the
coast-down test would not yield new information. We must
change the system in a quantifiable way. The most practical
approach is to couple to the motors shaft a known mass moment of inertia, a load JL, supported on a shaft with bearings
with unknown damping bL, Fig.4.88. When we repeat the
test, the time constant of the modified system is

JM + JL
bM + b L

4 Mechanical Systems

244
Fig. 4.88 Load inertia, JL, attached to the motors shaft with
a flexible coupling. The load
inertias shaft is supported on
bearings with damping bL

Motor bearings
damping bM

Load mass moment


of inertia J L

Motors rotor mass


moment of inertia JM

Flexible coupling

TC
g

Load bearings
damping bL

We know JL. We can set an upper limit for the load inertias
damping bL with the reasonable assumption that the load inertias damping is not greater than the motors damping, bL bM ,
and, for the initial curve fitting, that they are equal, bL = bM .
Manual calculations. The second approach is to approximate the derivative of the motors angular velocity in the
system equation by a finite difference,
d 1g
dt

d 1g
dt

Motor

100

Motor with Load

0
0

0.2

0.4

0.6

t, sec

0.8

1.0

Fig. 4.89 Locations for the finite difference approximations of coastdown test data

The motor coast-down data yields two useful finite


difference equations
1
TC = J M
+ bM 1g ( 0)
t1
and
2
TC = J M
+ bM 1g ( 0.15)
0
t2

1g
+ bM 1g TC J M
+ bM 1g
t

The very top and very bottom of the coast-down curve are
the most convenient segments to use for the finite difference equation, Fig.4.89. The angular velocity is a maximum at the top, so the effect of the velocity dependent,
viscous friction on the deceleration is greatest at the top
of the curve. At the very bottom of the curve, the angular
velocity is zero, so the viscous friction should have no effect on the final deceleration of the motor. Fit tangents to
the curve at the top and bottom of the data to estimate the
derivative of the angular velocity. Use the smallest value
t practicable to increase the precision of the finite difference approximation. The estimated derivatives will have
substantial uncertainty.

Estimate the slopes at


(0) and (t)=0

(t) 300
rad
___
sec 200

1g

The finite difference approximation must be made over


as small an increment of t as practicable to reasonably
approximate the derivative. If the measurement of angular
velocity were ideal, we would use t equal to one time step
in the data. However, there will be noise on the tachometer
signal. We may need to expand t slightly, so that velocity
values at the two end points of the interval do not appear to
be distorted by noise. Otherwise, it is essential that t be as
small as possible.
Having approximated the derivative with a finite difference, we can then evaluate the system equation since we also
know 1g
TC = J M

400

Since 4 g (0.15) = 0, the second equation simplifies to


2
TC = J M
t2
The coast-down test performed with the known load inertia
JL coupled to the motor shaft yields a second set of equations
3
TC = ( J M + J L )
+ (bM + bL ) 1g ( 0)
t3
and
4
TC = ( J M + J L )
+ (bM + bL ) 1g ( 0.90)
0
t4

4.7 Dynamic Tests

14.5

Model

2.0

Time Step
t = 110-4 sec

14.0

Data

245

= 0.1

13.5 Data Quantification


rad
___

Negative Error
Model - Data

= 0.6 sec

f(t)
___
1.0
f(ss)

rad 13.0
___
sec
12.5
12.0

0.2
0.3
0.4
0.5

1.5

0.201

t, sec

0.202

0.203

Fig. 4.90 Coast-down model plotted against experimental data. Note


the coarser quantification of the vertical axis of the data relative to the
size of the time step. A positive error occurs when the models value is
greater than the data

Equating the damping bL of the load bearings with the damping bM of the motor yields
3
TC = ( J M + J L )
+ 2bM 1g ( 0)
t3
and
4
TC = ( J M + J L )
t4
We now have four equations with three unknowns. The typical result of having more equations than we need and uncertain approximations of the finite differences is a range of
values for TC, JM, and bM.
The estimates of TC, JM, bM, and bL are refined by curve
fitting the coast-down models to the data. The curve fitting
can be performed either manually by iterating the parameter
values, until the fit between the model and the data is satisfactorily fit. One could also do so automatically by use of
a minimization routine in Mathcad or MATLAB. Minimization calculations reduce error between the model and the
data, Fig.4.90. To prevent positive and negative errors from
canceling each other, the absolute value of the error or the
square of the error is summed
 SumAbsoluteValueError = 0

N 1

SumErrorSquared = 0

N 1

Model p Data p

= 0.707

0.5

Positive Error
Positive Error Model > Data Model - Data
Negative Error Model < Data

0.200

= 0.9

(4.50)

(4.51)
Model p Data p

Mathcad has a minimization function, which is used inside
a solve block. MATLABs minimization function, fmincon(), is part of the add-on optimization toolbox.
A simple and surprisingly effective minimization
technique uses a genetic algorithm, in which sets of
parameters are randomly changed, retaining those which

0.5

1.0

t, sec

1.5

2.0

Fig. 4.91 A family of underdamped second-order step responses normalized to the steady-state value and parameterized by the damping
ratio

minimize the error function. A typical algorithm may have


10 or 20 sets of parameter values. The parameter values
tested in an iteration include (1) the parameter set which
minimized the error function in the previous iteration, (2)
parameter sets created by making random changes to the
that set, and (3) parameter sets created from random values
within broad limits. The algorithm is illustrated in Chap.8.
MATLABs optimization tool box also contains a genetic
algorithm function, ga().

4.7.4 Second-Order System Step Responses


The techniques characterizing a second-order systems step
responses, introduced in Sects.2.9.2 and 2.9.3, depend
on the damping ratio of the system and the type of input.
Overdamped systems have two real eigenvalues, s1 = 1
and s2 = 2, a damping ratio 1, and non-oscillatory impulse and step responses. The eigenvalues of underdamped,
oscillatory second-order systems, with <1, are a complex
conjugate pair, s1 , s2 = jw . Overdamped and underdamped second-order systems with damping ratios greater
than approximately 0.7 are more difficult to characterize
than less damped systems, because their impulse and step
responses oscillations which are distinctive and yield the
systems eigenvalues, Fig.4.91.
Characterization of second-order systems impulse and
step responses is easiest using Laplace transformations and
transfer functions, Sect.2.10 and Table2.3. The relevant
transform pairs of Table2.3 are reproduced as Table4.5, for
convenience. Note that the Laplace transformation of the
unit impulse is the number one.
A transfer function is a multiplicative linear operator. We
are familiar with multiplicative unit conversion. For example, the conversion from inches to centimeters is performed
by multiplying by the factor 2.54 cm/in, as shown below.

4 Mechanical Systems

246

1
1
1
=
s ( s + a )( s + b )
( s + a
)( s + b)
s

Input

Table 4.5 Selected Laplace transform pairs


f(t)

F(s)

(t ) = 0 for t 0
(t ) = for t = 0

Unit step: us (t ) = 0 for t < 0

1
s

us (t ) = 1 for t > 0

1
(e at e bt )
ba

( s + a )( s + b)
1
s ( s + a )( s + b )

1
1

1+
be at ae bt
ab a b

n
1

Transfer
Function

e nt sin n 1 2 t

n2
s + 2 n s + n2
2

Transfer functions are multiplicative operators but in the


Laplace-domain, Eq.4.52,
x in


2.54 cm
= y cm
in

Output ( s )
= Transfer Function ( s ) G ( s )
Input ( s )
Input ( s )

(4.52)

Output ( s )
= Input ( s ) G ( s ) = Output ( s )
Input ( s )

Input

( s + a
)( s + b)

Transfer
Function

( s + a
)( s + b)

Energetic Equations
F (t ) = Fb + FK + FM

Compatibility or Path Eq:


Element Eqs:
Energy Eqs:

Fb = bv1g

v1g = v1g
dFK
= Kv1g
dt

E sys = E K + E M

EK =

dv1g

FM = M
2
K

F
2K

dt

EM =

1
Mv12g
2

d 2 FK
K
b dFK K
F (t ) =
+
+
FK
2
M
M dt
M
dt

(3.33)

2
1 dF (t ) d v1g b dv1g K
=
+
+ v1g
M dt
M dt
M
dt 2

(3.34)

Output

The Laplace transformation of a unit step input is identical


to the Laplace transformation of the operation of integration
with respect to time. For example,
Fig. 3.31a Force source springmass-damper system of Fig.3.23.
b Linear graph, Fig.3.24b

The output could be either the step response of the system


represented by the transfer function, or the product of two
operators, integration with respect to time and the transfer
function. The former would transform to the time-domain as
a response function. The latter would transform to the timedomain as the integration of the system equation.
The ambiguity of the Laplace-domain is more than offset by its advantages, but it does require a change in many
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).
The example second-order system is the force sourcespring-mass-damper system of Sect.3.2.4.3, shown as an
energetic schematic and a linear 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.

Continuity or Node Eq:

There is an ambiguity in the Laplace-domain which results


from the form of the singularity inputs: the impulse, step and
ramp. Without knowledge of the history of a calculation, it is
impossible to identify a term as either a transfer function or
a transformed variable, a signal. The impulse input yields an
output, the Laplace transformation of the response function,
which is identical to the transfer function. For example,
1

Output

The objective is to determine the eigenvalues of the system


and, possibly, the magnitude of the input to the system, from

x,v

F(t)
g

g
Lubricating fluid
Damping b

F(t)
g

4.7 Dynamic Tests

247

an observed impulse or step response, in order to calculate


the values of unknown parameters of the system. There are
two approaches. When the eigenvalues are known, they are
then equated with the expressions for the eigenvalues formulated from the characteristic equation derived from the
system equation. Alternatively, features and values of the observed response are equated with the corresponding Laplacedomain expressions. In the case of underdamped, oscillatory
systems, the coefficients of the systems characteristic equation are equated with the coefficients of the corresponding
Laplace-domain expression, which is written in terms of the
damping ratio and the ideal, undamped natural frequency
n, which are calculated from the observed response.
When the eigenvalues are known, it is always possible
to solve for two unknown parameters by formulating the eigenvalues in terms of the systems parameters by solving the
characteristic equation with the quadratic formula. Using the
system equation, Eq.3.34, if one of the parameters is known,
then the other two can be calculated. For example, if M is
known, find b and K.

3.5
3.0
2.5

v(t)
2.0
cm
___
sec 1.5
1.0

t2

0.5
0.0

t, sec

10

12

Fig. 4.92
Example overdamped second-order impulse response
t
t
v (t ) = 10 e 0.67 e 1.67 . Times, t1, t2, and t3, indicate the locations
of data used in the characterization of the response

K
b
s1 s2 = 4
M
M

2
1 dF (t ) d v1g b dv1g K
b
K
=
+
+ v1g s 2 +
s+
=0
2
M dt
M dt
M
M
M
dt

t1

t3

K
b
( s1 s2 ) = 4
M
M
2

2
2
K b
M b
2
2
= ( s1 s2 ) K =
( s1 s2 )
4

4 M
M M

b
K
b
4
M
M
M
s1 , s2 =
2

(4.53)

Solve for b
2

s1 =

b
1 b
K
+
4
M
2M 2 M

s2 =

1 b
b
K

4
M
2M 2 M

and
2

s1 + s2
2
2

b
1 b
K
b
1 b
K
+

=
+
4
4
M 2M 2 M
M
2 M 2 M

s1 + s2 =

b
M

b = M ( s1 + s2 )
Solve for K

The alternative approach using the Laplace-domain expressions is often easier, as will be illustrated in the following
sections.

4.7.4.1 Second-Order System Impulse Responses


An impulse response begins and ends at zero. Impulse responses are created by either an impulsive input, such as a
hammer striking an elastic object, or by a step input acting
on a system, in which a system equation has the derivative of
the input as the only input term, i.e., Eq.3.34.
Overdamped Second-Order Impulse Response: An overdamped systems impulse response, Fig.4.92 and Eq.4.54,
is the superposition of two real exponential decays. The constant C is the magnitude of a possibly unknown input


f impulse (t ) =

(4.54)

To avoid the potential confusion between the smaller eigenvalue, which is more negative, and the larger eigenvalue,
which has the smaller magnitude, we will work with the
response function written in terms of time constants

vest (t ) = A e 1 e 2

s1 s2
2
2
b
1 b
1 b
K
K b
=
4
+

2 M 2 M
M
M 2 M 2 M

C
e at e bt
ba

where =

4 Mechanical Systems

248

3.5

Characterization begins by finding the larger, slower time


constant which dominates the decay portion of the response;
the faster exponential, with the smaller time constant, has,
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
of the response to estimate the first time constant, 1.
t1

ln v (t1 ) = ln ( A) +

t1

t2

t2

ln v (t2 ) = ln ( A) +

ln v (t1 ) ln v (t2 ) =


1 =

t1 + t2

ln v (t1 ) ln v (t2 )

t3
t3
v (t )
3
1

e = e 2
Aest

1 = 1.70

(4.55)

v (t1 )
t 1

=A

A = 9.28

(4.56)

To estimate the faster, smaller, time constant, select a data


pair near the beginning of the growth of the response. Both
exponentials in the response function contribute to the initial
portion of the response
t3
t3

1
v (t3 ) = A e e 2

t3

=e

t3

v (t3 )
A

t3

t3

= e

t, sec

10

t3
v t
(
t
3)
ln
e 1 = 3
A
2
est

t3

v (t ) t3
ln 3 e 1

Aest

2 = 0.64

(4.57)

The unknown input impulse magnitude C is


C
ba

1 1
C = A
2 1

C = A (b a )

C = 9.07

The function used to create the data and the estimated response function are plotted in Fig.4.93. The accuracy of the
estimated values would be improved by iteration or use of a
minimization routine, if necessary.
The eigenvalues are used with the solution formulated in
terms of the systems parameters, M, K, and b, Eq.4.53, to
solve for two if the third is known
s1 = 1 =

Move the known quantities to the left side


v (t3 )

2 =

A=

We now use the same data to estimate A




Logarithms of negative numbers do not exist. Multiply


both sides by negative one and then take the natural logarithm

t2

t1 + t2

t1

t
t
vest (t ) = 9.28 e 0.64 e 1.70

Subtracting the second equation from the first eliminates the


unknown magnitude A
ln v (t1 ) ln v (t2 ) = ln ( A) ln ( A) +

t
t
v (t ) = 10 e 0.67 e 1.67 and the estimate of the response function

ln v (t1 ) = ln ( A) + ln e 1 ln v (t2 ) = ln ( A) + ln e 1

-t
____

Fig. 4.93
Example overdamped second-order impulse response

yields
t1

-t
____

vest(t) = 9.28 e 1.70 - e 0.64

0.0

t2
ln v (t2 ) = ln A e 1

0.5

Taking the natural logarithm of both expressions


t1
ln v (t1 ) = ln A e 1

-t
____

1.0

v ( t 2 ) = A e 1

and

-t
____

2.5

v(t) 2.0
cm 1.5
___
sec

t2

v (t1 ) = A e 1

v(t) = 10 e 1.67 - e 0.67

3.0

s2 = 2 =

1
1

1
= 0.58
1.7

s1 =

s2 =

1
= 1.56
0.64

4.7 Dynamic Tests

249
n=1

t1

v(t) 1
m
___
sec

n=2 n=3
n=4
y1
y2
y4
y3

f(t)
___
1.0
f(ss)

t2

-1

0.2

t, sec

0.4

0.6

Fig. 4.94 Underdamped impulse response. Times, t1 and t2, indicate the
peaks used to calculate the damped period Td and the damping ratio

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 eigenvalues directly. An alternative approach is to determine the
damping ratio and the ideal, undamped natural frequency
n, from the response. The Laplace transform pair for a unit
impulse response of a second-order underdamped system is
written in terms of the damping ratio and the undamped natural frequency n

wn

w n t
2

e
sin w n 1 t
f (t ) =


1 2

wd

w n2
F (s) =

s 2 + 2w n s + w n2

f (t ) = C

wn
1

e w nt sin w n 1 2 t

1 y1
ln
n 1 yn
1 y1
4 2 +
ln
n 1 yn

4Td

6Td

Fig. 4.95 Peak numbering n and amplitudes yn, relative to steady-state


used with the log-decrement formula, Eq.4.53

The times and amplitudes of the peaks are


v(0.041) = 2.09 m/sec , and t2, v(0.251) = 0.26 m/sec

t1,

1 y1
ln
n 1 yn
1 y1
4 +
ln
n 1 yn

1 2.09
ln

2 1 0.26
1 2.09
4 +
ln

2 1 0.26

= 0.32

(4.58)

The damping ratio is calculated from the response with the


log-decrement formula, Eq.4.57 and Fig.4.95. Note that the
peak numbers begin at one and that the amplitude values yn
are relative to steady-state. The steady-state value of an impulse response is zero, so the amplitude values can be used
directly


2Td

t, damped periods

where the observed or damped frequency d of the response


is the factor multiply time t.
The general case has an additional coefficient, C, which
scales the response and carries units


(4.59)

The damped period, which is the observed or actual period,


is approximately equal to the duration between adjacent
peaks or troughs. The true damped period is calculated
using the zero crossings, but the likely error locating the
zero crossings is great enough to justify using the peak
values
Td = t2 t1 = 0.251 sec 0.041 sec

Td = 0.210 sec

The observed or actual angular frequency is the time required


to progress through a full cycle (circle) of 2 rad

wd =

2 rad
2 rad
rad
wd =
w d = 29.9
Td
0.210 sec
sec

The relationship between the ideal, undamped natural frequency, n, and the actual, damped frequency, d, is found
in the sine of the impulse response function:

wd = wn 1 2

4 Mechanical Systems

250

The ideal undamped natural frequency n is calculated from


the damped frequency d:

wn =

wd
1

wn =

rad
sec
0.32

29.9

w n = 31.6

rad
sec

The eigenvalues of an underdamped second-order system are:

1
L dF (t ) = L
M
dt

1
b
K

sF ( s ) = s 2 +
s + V1g ( s )

M
M
M
1
s
M
=
b
K
F (s)
s2 +
s+
M
M

V1g ( s )

or, written in terms of the damping coefficient, , and the


ideal, undamped natural frequency, n:

The coefficient C, the magnitude of the input which scales


the response function, is found by evaluating the impulse response function for a time t other than that of a zero crossing.
A time corresponding to a peak value will minimize error
due to signal noise. Evaluating the impulse response function for time t1 = 0.041sec
f (t ) =

wn
1

w n t

v ( 0.041 sec )

sin w n 1 t


wd

rad
31.6
sec e (0.32)(31.6)(0.041)sin 29.9 rad 0.041sec
=

sec
1 0.322

{ }

1
b
K
sF ( s ) = s 2V1g ( s ) +
sV1g ( s ) + V1g ( s )
M
M
M

s1 , s2 = jw d

(4.60)
s1 , s2 = w n jw n 1 2

d 2 v1g b
dv1g K
2 + L
+ L v1g
M
dt
dt M

The parameter values are calculated by equating the characteristic function, the denominator of the transfer function,
with the denominator of the Laplace-domain expression of
the Laplace transform pair.
1
s
w n2
M
= 2
b
K
s + 2w n s + w n2
s2 +
s+
M
M
s2 +

b
K
= s 2 + 2w n s + w n2
s+
M
M

and then equating coefficients of like powers of s


b
K
= 2w n and
= w n2
M
M

The ratio of the observed value over this value yields the
coefficient C:

We will again assume that the mass M is known. The spring


constant, K, is found from the ideal, undamped natural frequency, n,
K
w n2 =
K = M w n2
M

m
m
sec
C=
= 0.10
rad
rad
20.7
sec

The damping coefficient b is found from the product of the


damping ratio, , and the ideal, undamped natural frequency,
n,

v ( 0.041sec ) = 20.7

rad
sec

2.08

b
= 2w n b = 2 M w n
M

The response function is:


rad
31.6
m
sec e (0.32)(31.6)(t )sin 29.9 rad t
v (t ) = 0.1

rad 1 0.322
sec
v (t ) = 3.35

m
e
sec

10.1t
sec sin 29.9

rad
t
sec

Two of the three parameter values can be calculated if one


is known. The easiest method uses the transfer function derived from the system equation, Eq.3.34,

4.7.4.2 Second-Order System Step Responses


Second-order step responses contain more information than
impulse responses since they have either an initial value or
a steady-state value. For example, the steady-state force carried by the spring of the example spring-mass-damper system can be found from the system equation, Eq.3.33,
d 2 FK ( ss )
K
b dFK ( ss )
K
+
+
F0 us ( ss ) =
FK ( ss )
1
M
dt
M
M
dt 2
0
0
F0 = FK ( ss )

4.7 Dynamic Tests

251

The transient portion of a second-order step response is analyzed using the same techniques employed with a second-order impulse response, recognizing that the second-order step
response decays to its steady-state value rather than zero.
This fact is incorporated in the log-decrement formula. It
must be remembered when characterizing an underdamped
second-order step response.

v(ss)-v(t 1)
10

f step (t ) =

2
0

C
1

1+
be at ae bt
ab a b

(4.61)

which, when written in terms of the known steady-state


value and the unknown transient terms is
f step (t ) = f step ( ss ) + A1e at A2 e bt

(4.62)

where
f step ( ss ) =
A1=

C
ab

b
C
a b = a ( a b )

A2 =

t1 + t2

Fig. 4.96
Example
v (t )

t
= 10 16.7e 0.67

v ( ss ) =

overdamped

t
+ 6.7e1.67.

C
ab

C
A1 =
a ( a b)

b =

10

second-order

step

response,

v ( ss ) =

A1 =

1
1C

v ( ss )

11

1 1

C=

v ( ss )

A1 =

bv ( ss )
1

b
1

bv ( ss )
v ( ss )
v ( ss )
1
1

1 =

=
+1
A1
b1
A1
b1
A1

v ( ss )
1
= 2 = 1
+ 1
b
A1

2 = 0.64

Back-substitute to calculate the magnitude of the step input C

A1 = 16.2

The analysis of this overdamped step response now diverges


from that of an impulse response. An overdamped impulse
response has a single coefficient, whereas an overdamped

Times t1 and t2 indicate the locations

1 = 1.68

t, sec

step response has two. The steady-state value and the coefficients, A1 and A2, are expressed in terms of the magnitude
of the step input, C, and the inverses of the time constants,
a and b, in Eq.4.60, giving us three equations in three unknowns. We know the steady-state value and we have estimated 1 and A1, where

C=
= A1

of the data used to characterization of the response

C a
C
=

a
b
b
a

ab
( b)

We now use the same data and the estimate of 1 to estimate A1

t1

a=

ln v (t1 ) v ( ss ) ln v (t2 ) v ( ss )

v (t1 ) v ( ss )

C
ab

As with the impulse response, the characterization of an underdamped second-order step response begins with estimation of the larger, slower time constant. Two data pairs are
selected in the second half of the transient period, where the
response approaches but has not yet reached steady-state,
Fig. 4.96. The difference between the steady-state value,
v(ss), and the value of the response at time, t1, Eq.4.53, is
used to calculate the time constant

1 =

t1

v(t) 6
cm
___
sec 4

1
1

1+
f unit (t ) =
be at ae bt

ab a b
step

The general case for a step input of an unknown magnitude is

t2

Overdamped Second-Order Step Response: The Laplace


transform pair yields the unit step response function

v(ss)-v(t 2)

v( ss )

C=

1 2

v( ss )

C = 9.3

The coefficient A2 is:


A2 =

C
b ( a b)

A2 =

C
1 1 1

2 1 2

C = 5.93

4 Mechanical Systems

252

400

10
8

v(t)
cm
___
sec

300

(t)
rad 200
___
sec

4
2
Error: Initial values less than zero

0
-1

t, sec

10

Fig. 4.97 Example of an overdamped second-order step response


v (t )

t
= 10 16.7e 0.67

t
+ 6.7e1.67

100
0

0.5

1.0

1.5

t, sec

2.0

Fig. 4.98 Third-order systems step response

solid line, plotted against estimated


t

response function v (t ) = 10 16.2e1.68 + 5.9e 0.64 dotted line

f step (t ) = f step ( ss ) + A1e at A 2 e bt


t

vest (t ) = 10 16.2 e 1.68 ( 5.93) e 0.64


Note the estimated response function, vest(t), has an initial
error; its initial values are less than zero. The error between
an estimated response function and genuine experimental
data, as opposed to these computer generated data, is reduced though iteration, preferably with a curve fitting minimization routine, Fig.4.97.

4.7.5 Higher-Order System Responses


A higher-order system is a system above second-order. The
responses of higher-order linear systems can be constructed,
by superposing the responses of first and second-order systems. We will defer discussion of higher-order systems until
the introduction of state-space in Chap.7, because the reduction of the energetic equations of a higher-order system to a
system equation can be time consuming. However, we will
illustrate that higher-order systems responses are comprised
of superposed first and second-order responses, by decomposing a third-order rotational systems step response,
Fig.4.98.
Decomposition is the term for reversing the process of superposition. Decomposition is accomplished by characterizing the lower-order responses which superpose to create the
higher-order response, creating the corresponding impulse
or step response function, and then subtracting it from the
higher-order response. This is only practical, if the higherorder response data are available as a computer file.
The step response data shown in Fig.4.98 is from the tachometer of the DC servomotor used in the coast-down tests,
Sect.4.7.3.3. The system was configured with the load in-

ertia separated from the motors shaft by a torsion spring,


Fig.4.99. The system was running in steady-state at an initial
angular velocity 0, when the voltage applied to the system
was given a step increase. The initial angular velocity was
subtracted from the data of Fig.4.98.
We shall see in Chap.6 how to couple the electrical and
mechanical subsystems into a single model, but for the time
being, we shall model the electromechanical system as a
mechanical system, by representing the electrical aspect of
the motor as a torque source. The system is a third-order
system, because the motors rotor mass moment of inertia and the load inertia can have different velocities, since
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,
when the voltage to the motor is changed, the motors inertia begins to change speed before the load inertia. The difference in speed between the motor and the load changes
the amount of twist in the torsion spring and, consequently,
the torque through the spring. The changed spring torque
acts to decelerate the motors inertia and accelerate the load
inertia. The variation in motor speed creates the oscillation
in the tachometer data.
We will characterize the first-order step response, and
then remove it from the third-order step response to reveal
the second-order impulse response. The Laplace transform
pair for a first-order unit step response from Table2.3 is
1

at
f (t ) = a 1 e

1
F (s) =
s (s + a)

Multiplying both the time-domain and the Laplace-domain


expressions by the factor C creates the general case of a step
input of any magnitude C

1st (t ) =

C
1 e at
a

1st ( s ) =

C
s (s + a)

4.7 Dynamic Tests

253

Fig. 4.99 Test system consisting


of the mechanical aspects of a
DC motor. The system is driven
by the torque T(t) created by
the electrical aspect of the motor,
the mass moment of inertia of
the motors rotor, JM, a torsion
spring with spring constant K,
and an inertial load, JL

Torsion spring
spring constant K

T(t)

g
Motor bearings
damping b M

Load mass moment


of inertia JL

Motors rotor mass


moment of inertia JM

Coupling

Coupling
Load bearings
damping bL

400

2
300

T(t)

bM JM

bL

JL

(t)
rad 200
___
sec

= 393
0.632 = 248

100

Fig. 4.100 Linear graph of the third-order mechanical subsystem


model

We need the time constant and the steady-state value to calculate the eigenvalue, s = a , and the magnitude factor, C. From
Fig.4.94, we estimate the steady-state angular velocity as

( ss ) = 393

rad
sec

We will work with the time constant where, if the eigenvalue is purely real then s= and
1

1
=
=
a

We have two methods for estimating the time constant .


Recall that a first-order step response progresses through
63.2% of its dynamic range during each time constant .
We can estimate the time constant by finding the time when
the system has progressed 63.2% of 393rad/sec. We need
to eyeball (visually approximate) the first-order step response curve due to the oscillation, if we use this method.
An alternative technique is to choose a point on the trace
half way between the peaks and troughs of the oscillation.
That point should lie on the first-order step response, and a
visual approximation of the curve is not needed. We would
then calculate the time constant using the response function,
since we know the value of the response and the time at each
point, Fig.4.100.

0.5
0.32 sec

1.0

1.5

t, sec

2.0

Fig. 4.101 Construction to estimate the time constant of the first-order


aspect of the third-order step response

We will estimate the time constant the time corresponding


to 63.2% progression through the dynamic range, ,
= ( ss ) ( 0) = 393

( ) = ( 0.632) 393

rad
rad
rad
0
= 393
sec
sec
sec

rad
rad
= 248
sec
sec

From Fig.4.101, the time constant is estimated to be


=0.32 sec

1st (t ) =

C
1 e at
a

1st (t ) = 393

1st (t ) = 1 e

rad
0.32
1
e

sec

We wish to subtract the value of the first-order step response


function first (t ) from the third-order step response data to
yield the second-order impulse response superposed on the
first-order step response. The data were acquired digitally at
a sampling period Tsample. We must rewrite first (t ) as a discrete time function using a time step t equal to the data ac-

4 Mechanical Systems

254

40

40
n=1

20

20

2nd(t)
rad
___
sec

2nd(t)

rad
___
sec

-20
-40

n=4

-20
0.2

0.4

0.6

t, sec

0.8

Fig. 4.102 Second-order impulse response decomposed from the


third-order response of Fig.4.94. If the impulse response was not symmetric about zero, then the estimated time constant would be iterated
to make it so

-40
0

first ( mt ) =

A
1 e a mt = 1 e
a

first ( mt ) = 393 1 e

m t
0.32

m t

second = m 393 1 e
m

1 y1
4 +
ln
n 1 yn
1 26.1
ln

4 1 9.4

1 26.1
4 2 +
ln

4 1 9.4

3Td = 0.22 sec

2 rad
rad
= 86.1
wd =
0.073sec
sec
The damping ratio is calculated by means of the log-decrement formula, Eq.4.59,

1 = 26.1

wn =

Td = 0.073 sec

rad
rad
and 4 = 9.4
sec
sec

0.4

= 0.054

Calculate the ideal, undamped natural frequency n from the


observed, damped natural frequency, d, and the damping
ratio ,

wd = wn 1 2

The subtraction yields the second-order impulse response


shown in Fig.4.102. The time scale is expanded to show the
initial portion of the response in Fig.4.103.
The damped frequency, d, is calculated from the damper
period, Td.

0.3

We now use the counter variable m to perform the subtraction


m t
0.32

0.2

t, sec

1 y1
ln
n 1 yn

The first-order step response model written in discrete time is

0.1

Fig. 4.103 Construction to determine the damped period Td and the


damping ratio of the second-order impulse response

quisition systems sampling period Tsample, so that the values


calculated from the model are synchronized with the data.
Otherwise, the subtraction of the first-order model from the
third-order data will yield nonsense. Define the integer counter variable m such that
mmax t = tmax

3Td

rad
sec

86.1

1 0.054

wn =

= 86.2

wd
1 2
rad
sec

We check the estimated damping ratio and ideal, undamped


natural frequency n by plotting an oscillatory second-order
impulse response against the data. The oscillatory secondorder unit impulse response is
f unit

impulse

(t ) =

wn
1

e w n t sin w n 1 2 t

Our undamped, natural frequency of 86.1rad/sec is much


larger than the maximum peak value of the observed impulse
response, so we must attenuate the unit impulse response.
We will include a magnitude factor, B, in the numerator to
scale the response. The impulse function, including the scaling factor B, is

impulse (t ) = B

e w n t sin w n 1 2 t
w
1 2
d

wn

4.8 Equivalent Elements

255

40

the relevant equivalent behavior of an equivalent element is


equal energy storage or dissipation. An equivalent element
must have the same power flow as the combined power
flows of the original elements to be energetically equivalent

20

2nd(t)
rad
___
sec

POriginal + POriginal = PEquivalent


Element A

-20
-40

0.2

0.4

0.6

t, sec

0.8

Fig. 4.104 Second-order impulse model, dotted line, plotted against


the data, solid line

We guess an initial value for the scaling factor, B=0.5. We


then plot the impulse response against the data, expecting
the need to adjust the scaling factor, B, and discover that,
excluding the first peak, we have an excellent fit to the data,
Fig.4.104.
Having validated the first and second-order models by fitting the data, we now calculate the eigenvalues of the system
from the time constant, , of the first-order model and the
damping ratio, , and ideal, undamped natural frequency, n,
of the second-order model,
s1 =
and

s1 =

1
s1 = 3.13
0.32

s2 , s3 = w n jw n 1

series
elements

F (v12 + v23 ) = Psum

series
elements

We often wish to simplify a system by replacing two similar elements which are in parallel or in series with a single
equivalent element. What does equivalent mean in this
context? We are dealing with energetic systems. Therefore,

Fv12 + Fv23 = Psum

series
elements

Fv13 = Psum

series
elements

Conversely, elements in parallel have the same across variable, Fig. 4.107. Again summing the power dissipated in the
dampers in parallel and the power dissipated in the single
damper in the equivalent system

parallel
elements

4.8 Equivalent Elements

FA v12 + FB v23 = Psum

FA v12 + FB v12 = Psum

s2 , s3 = 4.65 j86.1

Element

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
variables in the original elements, Figs.4.105 and 4.106.
Elements in series have the same through variable,
Fig. 4.108. Using F and v as generic through and across variables, and equating power dissipated in the two dampers in
series in the original system with the single damper in the
equivalent system yields

s2 , s3 = ( 0.054)(86.2) j86.2 1 0.0542

Fig. 4.105a Linear graph of


a mass damper system with
dampers, b1 and b2, in series. b
Linear graph of the equivalent
system with a single equivalent
damper in place series dampers,
b1 and b2

Element B

( FA + FB ) v12 = Psum

parallel
elements

The values of the equivalent parameters are determined by


writing the continuity, compatibility, and elemental equations for the portion of the linear graph, which contains
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 equation written in terms of the through and across variables for
the equivalent element. Although you will likely remember
some equivalent element equations, be careful if you wish to
generalize the parallel and series relationships. The equivalent parameter expressions depend on both these conditions: whether the elements are in parallel or in series; and

b1

b2

F(t)
g

b equiv

F(t)
g

4 Mechanical Systems

256
Fig. 4.106a Linear graph of
a mass-damper system with
damper, b1 and b2, in parallel. b
Linear graph of the equivalent
system with a single equivalent
damper in place of parallel dampers, b1 and b2

b1

b2

F(t)

b equiv

F(t)

g
Fig. 4.107a Linear graph of
dampers, b1 and b2, in parallel
between nodes 1 and 2. b Equivalent single damper replacing the
two parallel dampers between
nodes 1 and 2, bequiv = b1 + b 2

b1
1

parallel

bequiv
1

b2

Fig. 4.108a Linear graph of


dampers, b1 and b2, in series
between nodes 1 and 3. b
Equivalent single damper,
b1b 2
1 1
bequiv = +
=
b1 b 2 b1 + b 2
series

b1

whether the parameter is multiplying the across variable or


the through variable.
Translational and rotational mechanical elements combine identically, but the terminology describing their configuration differs slightly. Rotational dampers or rotational
(torsion) springs which have the same angular velocity nodes
are said to be coaxial rather than in parallel. The calculations
of equivalent translational dampers and springs below also
apply to the rotational analog.

Elements in parallel have the same across variable. The sum


of parallel elements through variables equals the through
variable of the equivalent element, Fig. 4.107.
Continuity: Fb1 + Fb2 = Fbequiv
Compatibility: v12 = v12
Elemental: Fb1 = b1v12

Fb2 = b2 v12

Fequiv = bequiv v12

bequiv

4.8.2 Dampers in Series


Elements in series have the same through variable, Fig. 4.109.
The sum of their across variables equals the across variable
of the equivalent element, Fig.4.108.
Continuity:

Fb1 = Fb2 = Fbequiv

Compatibility:
Elemental:

4.8.1 Dampers in Parallel

b2

v12 + v23 = v13

Fb1 = b1v12

Fb2 = b2 v23

Fbequiv = bequiv v13

Reduction:
v12 + v23 = v13

Fb1
b1

Fb 2
b2

= v13

Fb

equiv

b1

Fb

equiv

b2

= v13

1 1
+ Fbequiv = v13
b1 b 2 series

bequiv
series

Reduction:
Fb1 + Fb 2 = Fbequiv

b1v12 + b 2 v12 = Fbequiv

Fbequiv = b1 + b 2 v12

bequiv
parallel

Algebra with a sum of ratios often leads to error. It is often


clearer and less error-prone to avoid dividing by a sum of
ratios. We will use both methods to illustrate. First, we will
divide by a sum of ratios and then clear improper ratios

4.8 Equivalent Elements

257

Fig. 4.109a Two springs in


parallel between nodes 1 and
2. b Equivalent single spring
replacing the two springs in
parallel between nodes 1 and 2

K equiv

K1

= K1 + K 2

parallel

equiv
series

b1b 2

b1b 2 1 1
+
b1 b 2
Fb

equiv
series

v13

b1b 2
b1 + b 2

K1

1 1
+ Fbequiv = v13
b1 b 2 series

1 1
+ Fbequiv = v13
b1 b 2

b1b 2

Fb

= v13

equiv

Fb

equiv

Fb

equiv
series

Fb

equiv
series

1
v
1 1 13
+
b1 b 2
b1b 2

b1 b 2
b1

v13

equiv
series

b1b 2
b1 + b 2

b1 + b 2

b1 + b 2

b1b 2

b1 b 2

v13

v13

bequiv =

Elemental:

b2

dFK1

dFK2

= K1v12

dt
dFKequiv
dt

dt

= K 2 v12

= K equiv v12

Reduction:

b1 + b 2

equiv

K equiv

Compatibility: v12 = v12

b1b 2

Fb

K2

Continuity: FK1 + FK2 = FKequiv

b2
b1
+

Fb = v13
b1b 2 b1b 2 equiv

b1b 2

Repeating the calculation, we can avoid division by combining the sum of ratios as a single ratio over a common
denominator and then multiply both sides by the inverse
ratio

b1 + b 2

K equiv

K2

Fig. 4.110a Springs, K1 and K2,


in series between nodes 1 and
3. b Equivalent single spring
replacing the two springs in
series between nodes 1 and 3
K1K 2
K equiv =
K
series
1 + K2

Fb

b1b 2
b1 + b 2

FK1 + FK2 = FKequiv

dFK1

K1v12 + K 2 v12 = K equiv v12

dt

dFK2
dt

dFKequiv
dt

+ K 2 ) v12 = K equiv
(K1 


parallel

K equiv

parallel

v12

parallel

4.8.4 Springs in Series

v13

b1b 2
b1 + b 2

4.8.3 Springs in Parallel


Elements in parallel have the same across variable,
Fig. 4.109. The sum of their through variables equals the
through variable of the equivalent element.

Elements in series have the same through variable. The sum


of their across variables equals the across variable of the
equivalent element, Fig.4.110.
Continuity: FK1 = FK2 = FKequiv
Compatibility: v12 + v23 = v13
Elemental:

dFK1

= K1v12

dt
dFKequiv
dt

dFK2
dt

= K equiv v13
series

= K 2 v23

4 Mechanical Systems

258

Reduction:
v13 = v12 + v23

v13 =

1 dFKequiv
1 dFKequiv
v13 =
+
K1 dt
K 2 dt
dFKequiv
dt

1 dFK1
1 dFK2
+
K1 dt
K 2 dt

1
1 dFKequiv
v13 =
+
K1 K 2 dt
1
v
1
1 13
+
K1 K 2



K equiv

M1

M2

M equiv

Fig. 4.111a Masses in parallel. b Equivalent single mass,


M equiv = M1 + M 2

series

Clear fractions
KK
K1 K 2
K1 K 2
1
= 1 2
=
=
1
1
1
1
K1 K 2
K1 + K 2
K1 K 2 K1 K 2
+
+
+
K1 K 2
K1 K 2
K2
K1
1

dFKequiv
dt

1
1
+
K
K
12

v13

dFKequiv
dt

K1 K 2
v13
=
K
K
+
1
2

K equiv

K equiv

series

M1

Impossible
2

M2

series

series

K equiv

Fig. 4.112 Mass and mass


moment of inertia elements can
never be in series, since every
mass or inertia must be referenced to ground

K1 K 2
=
K1 + K 2

4.8.5Equivalent Mass and Mass Moment


of Inertia
The linear graph system incorporates inertia elements, translational masses, or rotational mass moments of inertia, as
special cases. They are, of course, very common energetic
elements. They are special cases in that they are not the
general case of a linear graph element. There are two restrictions which apply to inertia elements. First, a branch representing either a force accelerating a mass must be connected
(referenced) to ground. Second, there is no reaction force
from ground. The lack of a reaction from the ground node
is why the portion of the symbol connected to ground is a
dashed line rather than a solid line. Although a force acts
through a spring or a damper, a force does not act through a
mass to the ground node.
Mass and mass moment of inertia elements must either
be in parallel, sharing a velocity node, or separated by a
spring or damping element. Mass and inertia elements are

rigid. The element can have no compliance, or it would also


store strain energy or dissipate energy. Consequently, if two
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.
Conversely, if the two masses are separated by a spring
or damper then they cannot be combined into a single mass,
since they have independent velocities. The test for independence between power variables is whether we can imagine
an instant in which the power variable has a value of zero
in one element and a non-zero value. We choose an instant,
because we do not expect a system to remain in this state.
We choose zero as one of the values, because this eliminates
the possibility that power variables are different but linearly
related, as we shall is the case when levers, belt drives, and
gear sets are introduced in Chap.6.
The only way to combine masses separated by a spring
or a mass is to revise the model, by removing the element
separating the masses. Revising the model to simplify it is
justified in the case that the velocity difference between the
two masses is judged to be insignificant, or, equivalently, the
amount of energy stored or dissipated by the elements separating the two masses is judged to be insignificant.

Problems

259

Summary
Table 4.6 Translational mechanical systems E = Fx P = Fv
Energetic Attribute Element Eqs.
Energy Eqs.
Customary
Power variable
form
form
Mass
Kinetic energy
storage

dv1g

F = ma

Fm = m

F = K x
F = K x12

dFK
= Kv12
dt

dt

Em =

1 2
mv1g
2

EK =

FK2
2K

Spring
Strain energy
storage
Damper
Ideal viscous
friction

d x12
Fb = bv12
dt
Note use of c
F=c

Table 4.7 Rotational mechanical systems E = T P = T


Energetic Attribute Element Eqs.
Energy Eqs.
Customary
Power variable
form
form
Mass
Kinetic energy
storage

d 1g

T = I

TJ = J

T = K
T = K12

dTK
= K 12
dt

dt

EJ =

1
J 12g
2

EK =

TK2
2K

Spring
Strain energy
storage
Damper
Ideal viscous
friction

Tb = b 12

Problems
Reminders
1. Draw a linear graph by identifying nodes of distinct values of the across variable, velocity for mechanical systems, on the schematic. A rigid object has a single velocity. Ideal rods, bars, and shafts are rigid and massless.
Identify the two nodes of the across variable associated
with each element in the schematic by identifying nodes
on either end of the schematic symbol, except for translational masses and rotational inertias, Tables 4.6 and 4.7.
Mass and inertia are rigid. They have a velocity node
and an inertial (ground) reference. After identifying two
nodes for each element, eliminate the redundant nodes,
and then label the distinct nodes with numbers, except
ground, which is identified with a g. All ground nodes
are the same velocity.
2. Use nodal notation for the across variable drop and indicate the positive direction of the through variable in the

direction of the drop in the across variable, except for


sources. Sources increase the across variable in the direction of the through variable flow.
3. Check the units of the system equation in terms of the
power variables of the system and time, not fundamental
units or conventional units.
4. Use superposition to create a pulse input from scaled and
time shifted Heaviside step functions. Use superposition to create the response function, by scaling and timeshifting the unit step response with the same factors and
time shifts used to create the input function. Each step response in the response function must be multiplied by the
corresponding time shifted Heaviside unit step function
to zero-out the response of that term unit its corresponding input acts on the system.
5. A dynamic system has only one characteristic equation, independent of our choice of the output variable.
Consequently, there is only one time constant of a firstorder system. All of the power variables in the system
vary at the same rate, although their step responses
will differ. Some variables will grow, while others will
decay, but they do so in synchronization. Second-order
systems have two eigenvalues. If the eigenvalues are
real, then the system is overdamped and does not have
an oscillatory homogeneous or step response. If the eigenvalues are complex conjugates, then the system is
underdamped and the homogeneous and step responses
are oscillatory.
Problem 4.1 A rotational mechanical which consists of
a angular velocity source, a drag cup, with damping constant, 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 velocity 20rad/sec, applied at an unknown time, t < 0, before
a step input of (t)=50rad/secus(t) was applied at time,
t = 0 , as plotted in Fig. P4.1b.
4.1.a Derive the system equation that relates the applied
velocity input to
i The torque acting through the torsion spring.
ii The velocity drop across the drag cup.
iii The velocity drop across the torsion spring.
Check the units of the system equations in terms of the power
variables and time.
4.1.b The input velocity (t) shown in the plot. Determine
the unit step responses of the system equations derived
in part a. Use superposition to determine the responses
to the angular velocity input plotted in Fig.4.1b.
4.1.c Plot the responses using Mathcad or MATLAB.
Problem 4.2 A rotational mechanical system consisting of an angular velocity source, a drag cup (fluid coupling) with damping b and a flywheel with mass moment

4 Mechanical Systems

260
Fig. P4.1a System with an
angular velocity source, drag
cup (fluid coupling) and torsion
spring. b Angular velocity input
applied to the system

b
b

(t)

(t)

rad
___
(0+) = 30 sec

t, sec
rad
(0-) = -20 ___
sec

Fig. P4.2a System with an


angular velocity source, drag cup
(fluid coupling), and flywheel,
bAngular velocity input applied
to the system

Flywheel
Rotational Inertia J

(t)

rad
(0-) = 40 ___
sec

Fluid Coupling
Damping b

Fig. P4.3a Rotational mechanical system comprised of a torque


source, mass moment of inertia
J and damping b. b Torque input
applied to the system

T(t)

Hydrodynamic
Bearing

rad
(0+) = -20 ___
sec

T(t), N m

150

50
0

of inertia J is shown in Fig. P4.2a. The damping constant


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
a step input in angular velocity of 40rad/sec applied at an
unknown time, t < 0 , before a step input was applied at time,
t = 0, reducing the angular velocity input to 20rad/sec, as
shown in Fig. P4.2b.
4.2.a Derive the system equation that relates the applied
velocity input to
i The torque acting to accelerate the flywheel.
ii The angular velocity of the flywheel.
Check the units of the system equations in terms of the power
variables and time.
4.2.b Determine the unit step responses of the system equations derived in parta. Use superposition to solve the
system equations for the angular velocity input shown
in Fig. P4.2b.

t,sec

Angular Velocity
Input (t)

t, sec

4.2.c Plot the responses using Mathcad or MATLAB.


Problem 4.3 A rotational mechanical which consists of a
torque source T(t), a rotational inertia, J = 8 kg m 2, and a
hydrodynamic bearing with b = 5 N m sec/rad is shown in
Fig. P4.3a. The system had reached steady-state under the
torque of 50 N m applied at a unknown time, t < 0, when
at time, t=0, the input torque is increased to 150 N m,
as shown in Fig. P4.3b.
4.3.a Reduce the equation list to the differential system
equation which relates the input torque T(t) to
i The angular velocity of inertia J.
ii The torque TJ acting to accelerate the inertia J.
iii The torque acting shear the fluid of the hydrodynamic bearing, Tb.
Check the units of the system equations in terms of the power
variables and time.

Problems
Fig. P4.4 a System with an
angular velocity source, drag
cup (fluid coupling) and torsion
spring. b Angular velocity input
applied to the system

261

End of Shaft
Rigidly Attached
Compliant Shaft
Torsional Spring K

(t)

rad
(0+) = 40 ___
sec

Fluid Coupling
Damping b

Angular Velocity
Input (t)

Fig. P4.5a Rotational mechanical system comprised of a torque


source, mass moment of inertia,
J, and damping, b. b Torque
input applied to the system

t,sec

rad
(0-) = -20 ___
sec

Drag Cup Shaft rigidly


Damping b attached

Flywheel
Rotational Inertia J

T(t),N m

150
50
0

Input Torque
T(t)

t, sec

-100

4.3.b Solve the system equations of part 4.3a for the input
plotted in Fig. P4.3b.
4.3.c Plot the response using Mathcad or MATLAB.
Problem 4.4 A rotational mechanical system consisting of
an angular velocity source, a drag cup (fluid coupling) with
damping coefficient, b = 8 N m sec/rad , and a torsional
spring with spring constant, K = 60 N m/rad, is shown in
Fig. P4.4a. The system has reached steady-state under the
input of (t ) = 20 rad/sec applied at an unknown time,
t < 0, when at time, t = 0, the input angular velocity is
increased to (t ) = 30 rad/sec, as shown in Fig. P4.4b.
4.4.a Derive the system equation that relates the input angular velocity to
i The angular velocity of the across the spring, 2g.
ii The torque acting through the spring, TK.
iii The velocity drop across the drag cup (fluid coupling), 12.
Check the units of the system equations in terms of the power
variables and time.
4.4.b Determine the unit step response and the use superposition to solve the system equations of part a for the
velocity input (t) shown in Fig. P4.4b.
4.4.c Plot the response using Mathcad or MATLAB.

Problem 4.5 A rotational mechanical which consists of a


torque source T(t), a rotational inertia, J = 5 kg m 2, and a
hydrodynamic bearing with b = 8 N m sec/rad is shown in
Fig. P4.5a. The system had reached steady-state under the
torque of 50 N m applied at a unknown time, t < 0, when
at time, t = 0, the input torque pulse shown in Fig. P4.5b is
applied to the system.
4.5.a Reduce the equation list to the differential system
equation which relates the input torque T(t) to
i The angular velocity of inertia J.
ii The torque TJ acting to accelerate the inertia J.
iii The torque acting shear the fluid of the hydrodynamic bearing, Tb.
Check the units of the system equations in terms of the power
variables and time.
4.5.b Determine the unit step response and the use superposition to solve the system equations of part a for the
input shown in Fig. P4.5b.
4.5.c Plot the response using Mathcad or MATLAB.
Problem 4.6 A rotational mechanical system modeled
as consisting of an angular velocity source, two dampers,
b1 = 60 N m sec and b2 = 5 N m sec, and torsional
spring, K = 40 N m/rad, is shown schematically in Fig. P4.6a.

4 Mechanical Systems

262
Fig. P4.6a Angular velocity
source (t) is connected to the
input shaft of a fluid coupling with
damping, b1. The output shaft of
the fluid coupling connects to the
torsion spring, K, the end of which
is free to rotate on bearings with
damping, b2, b Angular velocity
applied to the system by the source
(t)

End of Shaft
Rigidly Attached

(t)
rad
___
sec

Fluid Bearing
Damping b 2
Compliant Shaft
Torsional Spring K

600
400
200

Fluid Coupling
Damping b1

Ideal massless
and rigid bar

t, sec

10

-400

Angular
Velocity
Input (t)

Fig. P4.7a Force source F(t)


acts the bar which is attached to
dampers, b1 and b2. Damper b1
is connected directly to ground.
Damper b2 is connected to spring
K which is attached to ground.
b Force pulse applied to the
system

800

x,v

b1

F(t)
K

b2

F(t)
kN 2
0
-2

4.6.a Derive the system equations for


i The difference in the angular velocity of the two
ends of spring K.
ii The torque in damper b1.
iii The angular velocity across damper b2.
and check their units.
A step input of angular velocity (t ) = 20us (t ) was applied to the system at a time, t < 0 , sufficient for the system
to reach steady-state prior to when a second step input is applied at time, t = 0, as plotted in Fig. P4.6b.
4.6.b Solve the system equations.
4.6.c Plot the responses using Mathcad or MATLAB
4.6.d Plot the power flow from the source using Mathcad or
MATLAB
Problem 4.7 A translational mechanical system is modeled as consisting of a force source, two dampers,
b1 = 5.0 kN sec /m and b2 = 3.0 kN sec/ m , and a spring,
K = 4.0 kN/m, is shown in the schematic, Fig. P4.7a. The
bar is modeled as rigid, massless, and constrained to horizontal motion.
4.7.a Derive the system equations for
i The force in the spring.
ii The force in damper b1.
iii The velocity of the bar.

0.5

1.0

1.5

t, sec

and check their units.


The system de-energized when the pulse input of force
shown in Fig. P4.7b is applied at time, t = 0.
4.7.b Solve the system equations.
4.7.c Plot the responses using Mathcad or MATLAB.
4.7.d Plot the power flow from the source using Mathcad or
MATLAB
Problem 4.8 A rotational mechanical system modeled
as consisting of an angular velocity source, two dampers,
b1 = 60 N m sec and b2 = 5 N m sec, and a rotational inertia, J = 12 kg m 2, is shown schematically in Fig. P4.8a.
4.8.a Derive the system equations for
i The angular velocity of the inertia.
ii The torque in damper b1.
iii The torque in damper b2.
and check their units.
The system was in steady-state at time, t=0, under a
previously applied input when the angular velocity pulses
shown Fig. 4.8b were applied to the system.
4.8.b Solve the system equations.
4.8.c Plot the responses using Mathcad or MATLAB.
4.8.d Plot the power flow from the source using Mathcad
MATLAB.

Problems

263

Flywheel mass
moment of inerita J

Shaft rigidly attached


to support

Fluid coupling
damping b1
Angular
Velocity
Input (t)

Fluid coupling
damping b2

Fig. P4.8a Angular velocity source (t) is connected to the input


shaft of a fluid coupling with damping b1. The output shaft of the fluid
coupling connects to the mass moment of inertia J which is supported
on bearings with damping b2

(t)
rad
___
sec

800
600

200

-200

10

t, sec

12

-400
Fig. P4.8b Velocity pulses applied to the system by the angular velocity source (t)

Problem 4.9 A translational mechanical system is modeled as consisting of a velocity source, two dampers,
b1 = 5.0 N sec /m and b2 = 3.0 N sec /m , a spring,
K = 4.0 N/m , and a rigid, massless bar (a force spreader) is
shown in Fig. P4.9a. The bar is constrained to horizontal translation.

Fig. P4.9a Velocity source v(t)


acts on damper b1 which is attached to the force spreader bar.
Attached between the bar and
ground are spring K and damper
b2. b Velocity pulse applied to
the system

4.9.a Derive the system equations for


i The force in the spring.
ii The force in damper b1.
iii The velocity of the force spreader bar.
and check their units.
The system was at rest and relaxed before the velocity
pulse plotted in Fig. P4.9b acted on the system.
4.9.b Solve the system equations.
4.9.c Plot the responses using Mathcad or MATLAB.
4.9.d Plot the power flow from the source using Mathcad or
MATLAB
Problem 4.10 The translational mechanical system shown in
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,
b2 = 10 N sec /m , and M = 100 kg.
4.10.a Derive the system equations for
i The force from the velocity source.
ii The velocity of the mass.
iii The force acting to accelerate the mass.
and check their units.
The system was in steady-state under the previously applied step of 100m/sec at time, t=0 when the velocity input
plotted in Fig. P4.10b was applied to the system.
4.10.b Solve the system equations.
4.10.c Plot the responses using Mathcad or MATLAB.
Problem 4.11 A rotational mechanical system is shown in
Fig. P4.11a.
4.11.a Derive the system equations for
i The velocity of the flywheel J.
ii The torque acting through spring K.
iii The velocity difference across damper b.
and check their units.

Ideal massless
and rigid bar

v(t)

x,v

b
b2

4
2

b1
K

v(t)
m 0 0
___
sec
-2

-4

t, sec

4 Mechanical Systems

264

Fig. P4.10a Schematic of the


translational mechanical system,
b Velocity input applied to the
system

b
x,v

Lubricating
fluid film
Damping b2

b1

v(t)

m
v(t), ___
sec 3

2
1
0

10

-1

15

t, sec

-2

Fig. P4.11a A rotational system


consisting of torque source T(t),
a flywheel with mass moment
of inertia J, a compliant shaft
represented schematically as a
torsion spring K. The shaft turns
a drag cup with its output shaft
rigidly fixed. b Torque applied
by the source to the rotational
mechanical system

Shaft rigidly attached


Hydrodynamic
bearing damping b
Compliant shaft
torsion spring K

T(t),N m

Flywheel
rotational inertia J

100

The system was in steady-state under the torque input of


200Nm before the torque was increased to +200Nm at
time, t=0, and then to +400Nm at t=10sec, as shown in
Fig. P4.11b.
4.11.b Solve the system equations for the parameter values:

10

ii
iii

1
5

Nmsec
rad

300
200

Torque
input T(t)

b,

400

K,

N m
rad
1,500

J , kg m 2
25

1,500
3,000

30
10

4.11.c Plot the responses using Mathcad or MATLAB.


4.11.d Plot the power flow from the source using Mathcad
or MATLAB.
Problem 4.12 A rotational mechanical system is shown in
Fig. P4.12a.
4.12.a Derive the system equations for
i The torque acting through spring K.
ii The velocity difference across damper b1.
iii The velocity of the Flywheel J.
iv The torque acting to accelerate the flywheel.
and check their units.
The system was in steady-state under the angular velocity input of 100rad/sec before the velocity was changed to

-100

10

15 20

t, sec

25

+100rad/sec at time t=0, then to +400rad/sec at t=5 sec,


and then to 0 rad/sec at t = 10 sec. as shown in Fig. P4.12b.
4.12.b Solve the system equations for the parameter values:
b1 ,
i
ii
iii

N m sec
rad

1,000
100
1,000

b2 ,
0.5
0.5
0.5

N m sec
rad

Nm
rad
3,000
3,000
1,500
K,

J , kg m 2
50
50
500

4.12.c Plot the responses using Mathcad or MATLAB.


4.12.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
and M = 800 kg , and two dampers, b1 = 1, 000 N sec /m
and b2 = 2, 000 N sec /m , as shown schematically in Fig.
P4.13a.
4.13.a Derive the system equations for
i The force acting through damper b1.
ii The force acting through damper b2.
iii The velocity of mass M1.
iv The velocity of mass M2.
v The force acting to accelerate mass M2.
and check their units.

Problems
Fig. P4.12a A rotational system
consisting of velocity source
(t) driving a fluid coupling
with damping b1, a torsion spring
K, and a flywheel supported on
a bearing with damping b2. The
hydrodynamic bearing is represented schematically as a drag
cup. b Angular velocity applied
by the source (t) driving the
rotational mechanical system

265

Shaft rigidly attached


Hydrodynamic
bearing damping b2
Flywheel
rotational inertia J

b
(t) 400
rad 300
___
sec

Compliant shaft
torsion spring K

200

Fluid coupling
damping b1

100
2.5 5.0 7.5 10.0 12.5

Angular velocity
input (t)

t, sec

-200

Fig. P4.13a Force source F (t )


acts on a systems of two masses
and two dampers. b Force pulse
in kilonewtons applied to the
system

a
x,v

F(t)

b1

b2

M1

M2

2
1

F(t)
kN 0 0

t, sec
10

20

-1
-2
Fig. P4.14a Velocity source
v (t ) acts on a system of two
masses and two dampers,
b Velocity pulse applied to the
system

b
x,v

v(t)

Lubricating
fluid film
Damping b

2
1

v(t)
___
m 0 0
sec
-1

10

15

t, sec

-2

The system was at rest when the force pulse shown in


Fig. P4.13b was applied to the system. Note that the units of
force are kilonewtons.
4.13.b Solve the system equations.
4.13.c Plot the responses using Mathcad or MATLAB.
4.13.d Plot the power flow from the source using Mathcad
or MATLAB.
Problem 4.14 A translational mechanical system is shown
schematically in Fig. P4.14a. Parameter values are M=10kg,
b = 0.4 N sec /m , and K=3,000 N/m.
4.14.a Derive the system equations for
i The force acting through spring K.
ii The force acting through the lubricating film
damping b.
iii The velocity of mass M.
iv The force acting to accelerate the mass M.
and check their units.

The system was at rest when the velocity pulse shown in


Fig. P4.14b was applied to the system.
4.14.b Solve the system equations.
4.14.c Plot the responses using Mathcad or MATLAB.
4.14.d Plot the power flow from the source using Mathcad
or MATLAB.
Problem 4.15 A translational mechanical system is shown
schematically in Fig. P4.15a.
4.15.a Derive the system equations for
i The force acting through spring K.
ii The velocity drop across spring K.
iii The velocity drop across damper b.
iv The velocity of mass M.
and check their units.
The system was at rest when the velocity pulse shown in
Fig. P4.15b was applied to the system.

4 Mechanical Systems

266
Fig. P4.15a Velocity source
v (t ) acts on a systems of two
masses and two dampers. b Velocity pulse applied to the system

b
x,v

v(t)

2
1

v(t)
___
m 0 0
sec

10

-1

15

t, sec

-2

Youngs modulus E
Density
Width b

M
h

Fig. P4.16 A cantilevered beam with a cylindrical mass attached at


distance L from the support

4.15.b Solve the system equations for the parameter values:


i M=500kg, b = 2, 000 N sec /m, K=1,000N/m.
ii M=300 kg, b = 400 N sec /m , K=4,000 N/m.
iii M=500kg, b = 1, 000 N sec /m , K=6,000N/m.
4.15.c Plot the responses using Mathcad or MATLAB.
4.15.d Plot the power flow from the source using Mathcad
or MATLAB.
Fig. P4.17a Rotational
mechanical system. b Plan and
cross-section of the flywheel

Problem 4.16 A cantilevered beam with rectangular crosssection is shown in Fig. P4.16. Attached to the beam is a
cylinder of mass M. The diameter of the cylinder equals the
width of the beam b. The height of the cylinder is twice its
diameter.
4.16.a Determine the undamped angular frequency of the
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
and L=50 mm, b=10 mm, and h=1 mm.
4.16.c Determine the undamped angular frequency of the
beam and mass system when they are aluminum
6061 and L=50 mm, b=10 mm, and h=1 mm.
Problem 4.17 A rotational mechanical system consisting
of a torque source driving a shaft with a flywheel is shown
in Fig. P4.17. The shaft and flywheel are carbon steel. The
shaft is solid and has a circular cross-section. It is sup

Mass moment of inertia, J


Rolling-contact bearing, b

T(t)

b
A

f
a
b
c
d

Plan

Section A-A

Problems

267

Table P4.17 Flywheel and shaft dimensions


i
ii
iii

a (in.)
1.5
2
3

b (in.)
3
4
5

c (in.)
18
20
30

d (in.)
20
24
36

e (in.)
0.5
0.5
0.5

f (in.)
2
3
4

L (ft.)
10
20
30

ported by three rolling-contact bearings, each with damping


b = 0.5 N m sec/rad.
4.17.a Calculate the torsion spring constant K of the shaft
and the mass moment of inertia J of the flywheel for
the following dimensions.
4.17. b Determine the angular velocity of the flywheel when
a step input of 200N-m of torque is applied to the
de-energized system.
4.17.c Plot the response of the system in Mathcad or MATLAB.

Fig. P4.18a Translational


mechanical system, b Velocity of
the ideal massless and rigid bar

Problem 4.18 A test was performed to determine the energetic element parameters of translational mechanical system
shown in Fig. P4.18a. The system was de-energized before
it was subjected to a step input of force, F (t ) = 100.0 N us (t ).
The velocity of its point of application of the force on the
ideal massless and rigid bar was measured. The data are presented in Fig. P4.18b and Table P4.18.
Determine the damping coefficient b and the spring constant K. Report your results in the correct SI units.
Problem 4.19 A machines energetic attributes are modeled
as a mass moment of inertia J supported on an ideal, rotational damper b, Fig. P4.19a. A test was performed in which
a step input torque of 1,000N-m was applied to the system
and the angular velocity of the inertia measured, Fig. P4.19b
and Table P4.19. Determine the magnitudes of the inertia J
and the damping coefficient b. Interpolate if necessary. Report your results in the correct SI units.

x,v

K
F(t)

v(t)
m
___
sec

Time (sec)
0.0
0.5
1.0
1.5
2.0
2.5
3.0

Velocity
(m/sec)
4.000
3.023
2.285
1.727
1.305
0.986
0.745

2
1
0

Ideal massless
and rigid bar

Table P4.18 Velocity data

Time (sec)
3.5
4.0
4.5
5.0
3.5
5.5
6.0

Velocity
(m/sec)
0.563
0.426
0.322
0.243
0.563
0.184
0.139

Time (sec)
6.5
7.0
7.5
8.0
8.5
9.0
9.5
10.0

t, sec

Velocity
(m/sec)
0.105
0.079
0.060
0.045
0.034
0.026
0.020
0.015

10

4 Mechanical Systems

268
Fig. P4.19a Rotational
mechanical system, b Angular
velocity of the mass moment of
inertia J

b
T(t)

Hydrodynamic
Bearing

(t)
rad
____
sec

1.5

1.0

0.5

0.0

Table P4.19Angular
velocity data

t, sec

Time (sec)

1g(kRPM) Time (sec)

1g(kRPM) Time (sec)

1g(kRPM) Time (sec)

1g(kRPM)

0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
0.45
0.50
0.55
0.60

0.00
0.16
0.29
0.41
0.52
0.61
0.69
0.76
0.82
0.87
0.92
0.96
1.00

1.03
1.06
1.08
1.10
1.12
1.14
1.15
1.16
1.17
1.18
1.19
1.20
1.21

1.21
1.22
1.22
1.22
1.23
1.23
1.23
1.23
1.24
1.24
1.24
1.24
1.24

1.24
1.24
1.24
1.25
1.25
1.25
1.25
1.25
1.25
1.25
1.25

0.65
0.70
0.75
0.80
0.85
0.90
0.95
1.00
1.05
1.10
1.15
1.20
1.25

References and Suggested Reading


Boyce MC, Arruda EM (2000) Constitutive models of rubber elasticity:
a review. Rubber Chem Tech 73:504523
Budynas RG, Nisbett KJ (2011) Shigleys mechanical engineering
design, 9th edn. McGraw-Hill, New York
Ogata K (2003) System dynamics, 4th edn. Prentice-Hall, Englewood
Cliffs

1.30
1.35
1.40
1.45
1.50
1.55
1.60
1.65
1.70
1.75
1.80
1.85
1.90

1.95
2.00
2.05
2.10
2.15
2.20
2.25
2.30
2.35
2.40
2.45

Rowell D, Wormley DN (1997) System dynamics: an introduction.


Prentice-Hall, Upper Saddle River
Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
dynamics. Addison-Wesley, Reading
Timoshenko SP, Gere SM (1997) Mechanics of materials, 4th edn.
PWS-Kent, Boston

Fluid, Electrical, and Thermal Systems

Abstract

Fluid and electrical systems are important means of transmitting, transforming, and converting 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 potentialdriven flow, similar to low velocity (seepage) fluid flow.

5.1 Fluid Systems


Fluids, both liquids and gases, are used in machine design
to transmit power, store energy, and actuate mechanisms.
Fluids have mass and are compressible. Thus, they store
both kinetic and strain energy. Further, fluids are generally
confined within containers such as tubes, pipes, tanks, and
cylinders. An additional mode of energy storage is the work
that is done by those forces, which are exerted by fluids on
their containers. One important form of energy storage in
low pressure systems is the elevation of liquid against gravity. Mechanical engineers work with many different types of
fluid systems. Our emphasis will be on fluid power systems,
either pneumatic or hydraulic. These analyses, however, can
be applied to other types of fluid systems, such as water distribution and fire protection systems, as well.
Fluid power systems use both liquids and gases as the
working fluid. Hydraulic systems typically use oil as
the working fluid. They operate at pressures ranging from
750psi to 15,000psi. Fixed industrial equipment typically operate at pressures from 2,500 to 3,000psi. A pressure
of 5,500psi is common in mobile construction equipment.
Pneumatic pressures range from conventional low-pressure
pneumatic systems of 100150psi to high-pressure pneumatic systems with pressure up to 7,000 psi. Some pneumatic systems use single gases, such as nitrogen, rather than

the mixture of gases found in air. Vacuum systems are used


in automobile design, exploiting the difference in pressure
between the intake manifold and the atmosphere.
Water and wind power systems are ancient technologies.
The invention of steam power can be attributed to the Industrial Age. Hydraulic and pneumatic systems have long
been a means to apply force over a limited displacement,
with vehicle brakes being the most common example. A hydraulic piston/cylinder, a linear hydraulic motor, is also the
most familiar hydraulic component, since it is often visible.
Hydraulic systems have gained popularity in machine design, due to the diminishing cost of hydraulic components,
resulting from increased use of computer numerical control
(CNC) machine tools. Rotational hydraulic motors are now
used to eliminate shafts and gear sets (rotational mechanical power transmission elements), to reduce cost and increase design flexibility. High-pressure pneumatic systems
are unusual, but should become more common, if current
trends continue.

5.1.1 Fluid Power Variables


Mechanical engineers have an intuitive understanding of the
basics of fluid systems, because water flow is so familiar.
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

5 Fluid, Electrical, and Thermal Systems

270

Pipe, tube, or hose

Q(t)

Q(t)
p

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
Parabolic velocity profile

drop in the pipe. Knowledge of the pressure at one location


(node) is insufficient information to determine the direction
of flow. We must be able to calculate the difference in pressure between two nodes to determine the flow direction with
any accuracy.
Pressure drop in the direction of flow, p12, is the across
variable in fluid systems. We will measure flow as Volume Flow Rate, Q. Q is the through variable, which flows
through the pipe, Fig.5.1. Fluid power is the product of the
across variable times the through variable


P = p12 Q

(5.1)

where p12 is the pressure drop in the direction of the flow,


and Q is the volume flow rate.
We will simplify the flow of fluid through pipes, tubes,
and hoses by modeling the flow as slug flow, Fig.5.2. In
slug flow, the fluid velocity profile is uniform. The velocity of real fluid flowing down a pipe varies with radial and
longitudinal positions. Factors include the fluids properties,
the velocity of the flow, as well as the pipes geometry and
surface roughness. Assume the velocity is uniform across the
pipe, to visualize the fluid flowing down the pipe as a cylindrical slug, with velocity v. Although slug flow is a gross
simplification for some types of flows, it is a very accurate
approximation for turbulent flow, which is the typical flow
in fluid power systems. Turbulent flow occurs when fluid
passes through pumps with changes in pipe or tube diameters. Such flow also occurs at the sites of restrictions, such as
valves or the orifices of hydraulic cylinders. These components are spaced closely enough in typical fluid power systems to prevent the flow from ever developing the parabolic
velocity profile characteristic of laminar flow.
An assumption of slug flow (with its uniform fluid velocity profile) greatly simplifies calculation of volume flow
rates. The volume flow rate, Q, which passes through a
cross-section of the pipe with area, A, at time, t, is


Q(t ) = Av(t )

(5.2)

We will work in SI units, with volume flow rate expressed in


cubic meters per second.

Slug flow
Uniform velocity profile

Fig. 5.2a Parabolic velocity profile of fully developed laminar flow,


bUniform velocity profile of slug flow

m3
Q =
sec

(5.3)

A cubic meter is an awkwardly large volume to use in machine design. For a sense of scale, the mass of a cubic meter
of water is 1,000kg per metric ton. So, why use SI? In US
customary units, volume flow rate is expressed in units of
gallons per minute, abbreviated as GPM. However, US customary units for pressure, whether as pounds per square inch
(psi) or pounds per square foot (psf), are not derived from
gallons. One must convert gallons to either cubic inches or
cubic feet, for most calculations. For example, in the USA,
fluid power pumps are described by their flow rate in GPM
and outlet pressure in psi. In order to calculate a pumps
power in US customary units, one must convert gallons to
cubic feet and psi to psf, in order to yield power in units of
foot pounds per minute. The calculation is much more direct
in SI units


[ P ] = [ pQ ] = N2 m = Nm = Joule = Watt
m sec sec sec

(5.4)
SI also eliminates the need to calculate mass from weight.
We can dispense with having to remember the gravitational
constant in a variety of combinations of US customary units.
Finally, in the case of a hybrid system with an electrical subsystem, conversion to SI is required, because electrical units
of volts and amperes are SI.
Most mechanical engineers understand low-frequency
electrical systems by the fluid analogies. Fluid pressure is
analogous to electrical voltage. Fluid volume flow rate is
analogous to electrical current. The only significant difference between network representations of electrical and fluid
systems is that fluid capacitors must be referenced to atmospheric pressure, or some other ground pressure, whereas
electrical capacitors do not.

5.2 Fluid Elemental and Energy Equations

271

Pressure
Drop, p

Pressure
Drop p12

Operating Range
+/-p

Operating Point
Q 0, p0
Operating Range
+/-Q

p0

1
Q0

Volume Flow Rate, Q

Volume Flow Rate, Q R

Fig. 5.4 Linearization of a non-linear fluid resistance

Fig. 5.3 Linear fluid resistance, p12 = RQR

5.2 Fluid Elemental and Energy Equations


Fluid systems are mechanical systems, since they obey Newtons laws. They store energy as kinetic energy (energy of
motion) or strain energy (energy of elastic deformation).
Kinetic energy is dissipated as heat through viscous friction, which is lost from the system. One difference between
fluid systems and our treatment of translational mechanical
systems is that we will represent gravity as potential energy
storage in fluid systems, rather than as a force source.

5.2.1 Fluid Energy Dissipation: Fluid Resistance


Energy is dissipated as heat in fluid flow by viscous shear
in components, such as filters, orifices, and valves, due to
changes in diameters and direction of pipes and the roughness of their inner surface. In the simplest case, the pressure
drop in the direction of the flow is proportional to the flow
rate, Fig.5.3.


p12 = RQR

(5.5)

This relationship has the same form as electrical resistance,


v12 = RiR.
Unfortunately, the simplest case of fluid resistance does
not describe many real-world situations. It is an accurate
model for slow flows, such as flow of oil through a filter,
or water permeating through soil. The model loses accuracy,
however, as the fluid volume flow rate increases, because
fluid viscosity is not constant. Viscosity is a function of the
velocity of the fluid flow, more accurately described by a
power law viscosity function


( v ) = 0 v n

(5.6)

We must use linear elemental equations to produce linear


system equations. The simplest model for flow through a
fluid resistance is illustrated with a pressure drop across the
resistance that is linearly related to the volume flow rate, Q.

If we need to improve the accuracy of our model of fluid


resistance, we will linearize the model about an operating
point. The process, described in Sect.4.2.4.1, approximates
the pressurevolume flow rate curve by a straight-line tangent of the curve at the operating pointof the flow rate, Q0,
and pressure drop, p0, which is the expected operating condition of the system Fig.5.4. If the pressurevolume flow rate
curve is available as a mathematical function, the function is
linearized as a two-term Taylor series approximation, which
is equivalent to the graphical procedure.
The tangent to the operating point, Q0, p0, is


p12 = R QR + p0

(5.7)

where the term, QR, is a deviation above or below the operating point, Q0. Notice that this relationship cannot be inverted to calculate QR, since the expression does not contain
Q0. The expression yielding QR is

QR =

p12
R

+ Q0

(5.8)

5.2.2 Kinetic Energy Storage Fluid Inertance


Fluid has mass. Consequently, flowing fluid possesses kinetic energy. We will derive the Elemental and Energy equations for the dynamic element, which represents mass in a
fluid system, by expressing the corresponding equations for
a translational mechanical system in terms of the fluid power
variables, p and Q. Consider the mass of fluid with density
in a pipe of length L and cross-sectional area A, shown in
Fig.5.5. The cylindrical lump of fluid has mass M


M = AL

(5.9)

To accelerate or decelerate, a force must act on the mass of


fluid. Substituting M = AL for the mass of the fluid, and
expressing acceleration as the derivative of velocity with respect to time

5 Fluid, Electrical, and Thermal Systems

272

Density

F2 = p2 A
p

Fig. 5.6 Net pressure force acting on a fluid lump

Pressure

dv
F = AL
dt

= F1 F2 = Ap1 Ap2 = A( p1 p2 ) = Ap12

The power variable, Q, volume flow rate, is related to the assumed uniform velocity of the flow. The volume of fluid flowing at velocity v, that passes through a plane, normal to the
flow of area A, in a period of time, t, is volume V = Av t ,
Fig.5.7.
Hence, the volume flow rate, QI, is


Volume
= QI = Av
t

(5.10)

v=

QI
A

(5.11)

Now write F = ma, in terms of p12 and QI


FNet

Pr essure

= Ma

Q
d I
A
p12 A = AL
dt


p12 = I

dQI
dt

p12 =

L
A

(5.13)

E Kinetic =


EI =

1 A 2
QI
2 A
1 2
IQI
2

(5.14)

where fluid inertance I plays the role of mass in the equation.

p12 A = AL

I=

is called fluid inertance.


Note an apparent contradiction in the definition. Shouldnt
increasing the cross-sectional area of the pipe increase the
inertia of the fluid moving down it, by increasing the mass
of fluid in a length of pipe? It does because fluid inertance
I is not inertia. Fluid inertance is derived by using volume
flow rate QI in place of fluid velocity v. Increasing the crosssectional area of a pipe results in decrease in the fluid velocity for a given volume flow rate QI. The acceleration of the
fluid is also decreased proportionally. Hence, fluid inertance
I is inversely proportional to the cross-sectional area A of
the pipe.
Using the above expressions for the mass of fluid in a
length of pipe and the uniform velocity in terms of QI, we
can write the expression for the kinetic energy stored in a
flowing fluid

The velocity, v, expressed in terms of volume flow rate, Q, is




Fig. 5.7 Volume flow rate in slug flow

The force acting to accelerate the mass of fluid is the net


pressure force between node 1 and node 2, Fig.5.6,
FNet

This is the elemental expression for fluid inertia, the element


which stores kinetic energy in a fluid system. The quantity

dv
F=M
dt

Q(t)

Area A

M = AL

Volume

F = Ma

Fig. 5.5 Cylindrical fluid mass, a fluid lump

Velocity

Q(t)

Area A

F1 = p1 A

t2

t1

Length L

dv
dt

5.2.3Pressure-Based Energy Storage Fluid


Capacitance

L dQI
A dt

(5.12)

Fluid systems can store energy in the elastic deformation of


the pipe or structure containing the fluid under pressure, in
the elastic deformation of the fluid itself, and by raising fluid
against gravity. All three of these energy storage modes lead
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

pg

equations is given the analogous name, fluid capacitance,


and the same symbol, C, as electrical capacitance.
The elemental equation for a fluid capacitance relates the
volume flow rate into fluid capacitance, QC, to the rate of
change of pressure in the capacitance,


QC = C

dp2 g
dt

5.2.3.1 Gravitational Potential Energy


Gravity was not included as an energy storage mode in translational mechanical systems, because it only acts in one direction. It is easier to include gravity as a force source which
acts on mass. Gravity modeled as a force source still provides energy storage, because power can flow both into and
out of an ideal source. A car battery is a good example of
a source which can either deliver (source) or accept (sink)
power.
In fluid systems, gravity is represented as pressure-based
energy storage, where the pressure is created by the height
of a column of fluid. The most familiar example of such a
fluid capacitor is a municipal water tower. A water tower in
the form of a cylindrical tank without legs is called a standpipe.
The net pressure force acting on a slug of fluid to push
it into the standpipe results in mechanical work being done
on the fluid. The net pressure force acting on the fluid slug,
Fig.5.8, is

Pressure

p1g A slug

(5.15)

where C is the fluid capacitance. Note that pressure is measured relative to atmospheric or ground pressure in the elemental equation of a fluid capacitor. The voltage in an
electrical capacitor is measured across the capacitor. As capacitance C increases, more fluid volume in or out of it is
required to change pressure at a given rate.
In general, both the fluid and the components which contain the fluid are elastic and can store strain energy. These
strain energies are calculated independently. First consider
the fluid to be incompressible, and calculate the strain energy
stored in the components containing the fluid. Then, consider those components as rigid and calculate the strain energy
stored in the fluid. Gravitational potential energy storage is
usually insignificant in hydraulic systems in machines because the hydraulic pressures (on the order of 2,500psi) are
so much larger than the gravitational pressures. Gravitational
potential energy storage is important in water systems, both
within a building and in a municipal distribution system.

FNet

Area A
Density

= p1g p2 g Aslug = p12 Aslug

The mechanical work done to push the fluid slug into the
standpipe is

x,v

H
p2g A slug

p1

p2

Fig. 5.8 A standpipe-type fluid capacitance showing the work done


pushing a lump of fluid against the pressure in the capacitance

W
= FNet

Pressure

(5.16)

This can be expressed in terms of the power variables of a


fluid system, pressure, and flow rate. Differentiate the mechanical work:
dW
d
d
p12 Aslug x = p12 Aslug ( x ) = p12 Aslug v
P
=
dt
dt
dt

Then, use the definition of volume flow rate:


QC = Aslug v
We confirm that the rate of mechanical work being done to
push fluid into the fluid capacitance is fluid power:


dW
P
= p12 Aslug v = p12 QC
dt

(5.17)

Derive the fluid capacitance, C, of a standpipe. Assume the


fluid is incompressible, and the standpipe is rigid. The crosssectional area of the standpipe is A. The density of the fluid
is . Conservation of mass dictates the mass of fluid, which
flows into the port, is stored in the standpipe. Express mass
conservation in terms of mass flow rate:



Mass Flow Rate In =


QC = A

d ( MassStored )

dH
dt

dt

(5.18)
(5.19)

Introduce the power variable, pressure, by recognizing that


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:


p2 g = gH

(5.20)

5 Fluid, Electrical, and Thermal Systems

274

v2
QC

p1

x, v

p1g
p1g A

v3

FK

pg

Piston Area A

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
QC =


g
dH
A
g
dt

QC =

A
dH
g
g
dt

A dp2g
QC =
g dt

Pressure Cylinder

C=

A
g

Fig. 5.11 A bladder-type hydraulic accumulator. The compressibility


of the nitrogen gas in the bladder stores energy

A2 dp1g
QC =
K dt

QC
A

Hydraulic Lines

(5.22)

A free body diagram of the forces acting on the piston,


Figs. 5.9 and 5.10, leads to an equilibrium equation, which
relates the net pressure force, p1g A, of the fluid to the force
in the spring, FK.
The elemental equation for this fluid capacitor can be derived from the elemental equation for a spring.
dFK
= Kv23
dt

Low pressure
valve

(5.21)

5.2.3.2Energy Stored in the Elastic Strain


of the Fluid Container
Consider a fluid system component which consists of a
spring-loaded piston in a cylinder. This device is called a
fluid accumulatorin hydraulic systems, a pressure relief
tank in hot water systems, or a fluid capacitor in system
dynamics. Assume the fluid is incompressible, and the container is rigid, except for the spring. The volume of the cylinder which contains the spring is vented to the reference
(ground) pressure, usually atmospheric pressure. The only
force acting on the right side of the piston is the spring force,
FK. There are two different types of across-variable nodes
identified on the schematic, pressure, and velocity.
The volume flow rate, QC, is related to the velocity of the
piston relative to the end of the cylinder, v12.
v23 =

Bladder

Hydraulic
Oil

The coefficient term multiplying the time derivative of pressure fluid is the fluid capacitance,


N2 Gas

d p1g A
dt

)=KQ

(5.23)

The capacitance of a fluid accumulator is




C=

A2
K

(5.24)

The spring used in a fluid accumulator may be a conventional compression spring. An alternative design is to use a
diaphragm, rather than a piston, and high pressure gas, typically nitrogen, as the spring, Fig.5.11. A component may
act as an unintentional fluid capacitor, if it exhibits sufficient
elastic deformation when filled with a pressurized fluid. A
common example is a garden hose. Many garden hoses expand significantly when pressurized, causing a delay in delivery of water from the hose. A significant amount of strain
energy is stored in the hose, even under the relatively low
pressure of a residential water system of approximately
50psi. The energy is released when the tap is closed, allowing water to continue to flow out of the hose.
The energy equation for this fluid capacitor can be derived from the energy storage equation for the spring

EK =


FK2
2K

EK

( Ap )
=

1g

2K
1
2

EC = C p12g

EC =

1 A2 2
p1g
2 K
(5.25)

5.2 Fluid Elemental and Energy Equations

275

5.2.3.3Energy Stored in the Compression


of the Fluid
Positive stress is tensile in solids. The analog of stress in fluids is pressure where positive pressure is compressive. Liquids in fluid power systems are kept under positive pressure
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
flow from a region of negative pressure into positive pressure. Bubbles and dissolved air in hydraulic oil are always
problems because they greatly increase the compressibility
of the fluid.
Strain in a fluid is defined as volumetric strain, a fractional change in volume. Express volumetric strain by its effect
of increasing the density of the fluid, as a fractional change
in density


Volumetric Strain

dp =

(5.27)

where is the Bulk Modulus. This relationship has the


same form as Hookes Law for linear strain in a solid


= E

(1.3)

The similarity is stronger than it may appear, because


Hookes law can be written as
d = E

dL
L

To calculate the energy stored in the deformation of the fluid,


assume (1) the container is rigid, and (2) more fluid can be
packed into the fixed volume, Vol, of the container by compressing the fluid at increased pressure, Fig.5.12.
Conservation of mass requires that fluid flowing into a
container of fixed volume, Vol, must be stored in the tank, by
increasing the density of the fluid.
Mass Flow Rate In =

QC = Vol

d ( MassStored )
dt
d
dt

p1
pg
Rigid container, volume Vol
Compressible fluid, bulk modulus

Fig. 5.12 Increasing fluid pressure within a rigid container stores strain
energy in a compressible fluid

Expressing d in terms of the bulk modulus, , using


d
dp =
yields

(5.26)

In the simplest case, a linear relationship exists between the


applied pressure and the volumetric strain


QC

QC = Vol


dp1g
dt

Vol dp1g
QC =
dt

(5.28)

In this expression for fluid capacitance,




C=

Vol

(5.29)

5.2.4 Fluid System Sources


Sources are power supplies, in which one of the two power
variables of the system is either known or controlled. Fluid
system sources are either pressure sources or volume flow
rate sources. Pressure sources are more common than volume flow rate sources. Pressure sources occur naturally,
wherever there is a large reservoir at an elevated pressure,
relative to atmospheric pressure. Volume flow rate sources
do not occur naturally. Volume flow rate sources are created
by (1) placing a pressure source under feedback control to
vary pressure as the systems input resistance varies, or (2)
by placing a positive displacement mechanical pump under
velocity control. A positive displacement pumps geometry
displaces a certain volume of fluid every cycle. Pumps and
hydraulic motors are addressed in Chap.6.

5.2.4.1 Pressure Sources


Fluid pressure sources exist in nature and in technology.
A natural fluid pressure source is essentially a large standpipe-type fluid capacitance, that is, a fluid reservoir with

5 Fluid, Electrical, and Thermal Systems

276
Fig. 5.13a An industrial stationary fluid power unit driven by a
three-phase AC motor. The pump
is below the motor in the tank.
b A fluid power unit modeled
as a pressure source represented
schematically

Three phase AC motor


Breather cap. Filtered
vent to atmosphere

b
Vent to
atmosphere

P(t)

Pump

High pressure
Output from pump
Low pressure
Return to tank
Tank. Hydraulic
fluid reservoir

Stationary fluid power unit

sufficient height to create pressure, and sufficient volume for


that height change negligibly when there is flow out of or
into the reservoir. Earths atmosphere is a source of atmospheric pressure. An intake pipe in a lake or an ocean draws
water from a pressure source. Biological systems have various means of creating pressures or vacuums. Air and blood
flow from the pumping action of muscles. Water is drawn to
great height in trees, due largely to capillary tension.
Artificial pressure sources are either fluid capacitances or
pumps, as in liquids, and compressors, as in gases. A tank of
nitrogen gas, a fluid capacitance, is a pressure source. The
pressure source of a hydraulic automobile brake is a piston
in a cylinder. The pressure source for air-powered tools is an
air compressor. Industrial hydraulic systems are powered by
pumps. These devices are closer to ideal pressure sources,
when the flow rate drawn from them is small, relative to the
capacity of the device. There is finite resistance to any flow.
The greater the flow rate, the greater the effect of the fluid
resistances in the device, and the larger the pressure drop
within the source.
A stationary fluid power unit, as in Fig.5.13, provides
hydraulic power for industrial machinery. Water served as
the original hydraulic fluid, but its corrosive properties and
lack of lubricity (lubricating ability) led to its replacement
by mineral oils and emulsions of water and mineral oil. Hydraulic fluids most commonly used in industrial and mobile
equipment are mineral oils. A phosphate ester is used for
aircraft hydraulic systems due to the flammability of the hydraulic oil. Silicone-based hydraulic fluid is used for its nonflammability. An emulsion of water, mineral oil, and glycol
is another variety of hydraulic fluid.
The properties, including density and bulk modulus, vary
with the type of fluid and, importantly, with the amount of
air entrainment. The density and modulus of phosphate ester
and silicone-based hydraulic fluid are significantly lower
than that of oil-based hydraulic fluid. Although we use a
constant value for the bulk modulus, , in calculations, it is
actually a function of pressure. Various additives are mixed

High pressure
Discharge

Atmospheric
pressure Return

Fluid
Reservoir

Fluid power unit modeled


as pressure source p(t)

with hydraulic fluids for a number of purposes: to protect the


metal surfaces of the pumps and actuators in the system from
wear and corrosion; to limit foaming of the oil; to extend the
usable life of the fluid; and to change viscosity.
Air entrainment is a significant concern. Cavitation is
a phenomenon, whereby a local area of low pressure within a flow permits air to come out of solution, and form a
bubble. Extremely high stress results when such a bubble
collapses. Cavitation erodes hardened steel. The particle of
steel introduced into the hydraulic fluid, in turn, damages the
system. Hydraulic systems require filtration. The degree of
filtration, in terms of the size of particle removed, depends
on the nature of the system. Filtration for servo-hydraulic
systems with actuators under feedback control has the most
stringent requirements to protect servo valves and other precision components.
Foaming of hydraulic oil is also problematic. Foam floats
on the surface of a fluid in a reservoir and does not flow, if
the discharge (intake to the pump) is located below the surface, as one can observe in the shower.
Fluid power units are sized, in terms of the volume of
fluid delivered in a unit time at a specified pressure. US customary units for fluid power are gallons per minute and psi,
which are expressed inconsistently. Rather than convert gallons to cubic inches, it is best to convert both the volume and
the pressure to SI units, compute power in watts, and then
convert to the units needed to present the result.
The pump of a stationary fluid power unit is typically
powered by a three-phase AC motor. In order to maintain
pressure while the power unit is in stand by (not discharging fluid to the external system), there is a high-resistance
internal hydraulic bypass circuit from the pump to the fluid
reservoir, Fig.5.14. The reservoir is vented to the atmosphere through a filter in the fill cap, called a breather
cap.The filter is intended to keep foreign matter, especially
abrasive particles, out of the hydraulic fluid. The fluid reservoir is box shaped and sized so that the residence time,
the period of time the fluid stays in the reservoir during

5.3 Linear Graphs of Fluid Systems

Vent to
atmosphere

Q(t)

Pump

277
High pressure
Discharge

R internal

High resistance
internal by-pass

Fluid
Reservoir

Atmospheric
pressure Return

Fig. 5.14 Fluid power unit, modeled as a volume flow rate source. A
high-resistance internal bypass between the pump discharge and the
tank allows the pump to run, when the system does not accept flow

operation, is approximately two minutes. The residence


time allows the fluid to de-gas, i.e., have air bubbles rise
to the surface.
Mobile fluid power units have the same functional components. Power for the pump of a mobile unit is usually drawn
from the shaft of the internal combustion engine. Some electrically power pumps, however, are used in smaller systems.

5.2.4.2 Volume Flow Rate Sources


Volume flow rate sources do not exist in nature. There must
be technology involved, because the pressure to drive the
flow into the system must be varied to maintain a given flow
rate. Artificial volume flow rate sources accomplish this, by
varying the pump speed or the resistance of the internal bypass with feedback control, Sect.9.3.7.

5.3 Linear Graphs of Fluid Systems


The reduction of fluid systems to a system equation follows
the same procedure as for mechanical and electrical systems.
The first step is most important. Find the nodes of distinct
values of the across variable, pressure. There are always two
nodes associated with each element. Always put a node at
either end of pressure or flow sources, fluid resistors, and
fluid inertances. One node of a fluid capacitor must be atmospheric pressure (ground). The other node is the maximum
pressure in the fluid capacitor, which may occur anywhere
within a fluid accumulator, or at the bottom of a water tank
or standpipe.
Fluid system schematics contain ideal pipes, tubes, and
hoses, which have no energetic properties. Ideal pipes are
analogous to ideal conductors in electric circuits, as well as
massless and rigid rods or shafts in mechanical schematics.
Do not assign fluid element properties to an ideal pipe. If
a pipe, tube, or hose has energetic properties, they will be
indicated on the schematic. Every fluid element must have
two pressure nodes associated with it. Fluid resistance and
fluid inertances have pressure nodes at either end of the ele-

vent to
atmosphere

P(t)
Pump

R1
R2

Fluid
Reservoir

Fig. 5.15 Fluid system consisting of a pump modeled as a pressure


source, two fluid resistances, and a fluid capacitance

ment. Fluid capacitances contain one pressure node inside.


The other pressure node is the ground pressure, usually atmospheric pressure. The most common element to overlook
is a fluid inertance. Dont forget that F=ma applies to fluids.
There must be a pressure drop, to accelerate or decelerate
fluid flow in a pipe. Consequently, there are pressure nodes
at either end of a fluid inertance.

5.3.1 Example Fluid System Linear Graphs


5.3.1.1 Example One
The model of the fluid system shown in Fig.5.15 consists
of a pressure source which discharges into fluid resistance
R1. Fluid resistance R1 connects to the junction of a fluid accumulator (capacitance) C and fluid resistance R2. Fluid resistance R2 discharges into the pumps reservoir, completing
the fluid circuit.
Nodes of distinct pressure are identified in the schematic
Fig.5.16a. The ideal pipe between the discharge port of the
pump and fluid resistance is node 1, the discharge pressure
of the pump. The ideal pipes connecting resistances R1 and
R2 and the capacitance C are all node 2, the pressure of the
fluid capacitance. There is a pressure difference across resistance R2, but the pipe connecting R2 to the fluid reservoir is
ideal. Ideal pipes have fluid properties and, therefore, have
no pressure change along them. The fluid reservoir is vented
to the atmosphere. Therefore, the pressure in the ideal pipe
between the reservoir and resistance R2 is atmospheric pressure, g. Check for two nodes associated with each energetic element. Create the linear graph, by first drawing and labeling the nodes of distinct pressure. Then, add the branches,
which represent the elements between them. The branch for
a fluid capacitance is dashed after the arrowhead to indicate
that the fluid cannot flow to ground through the capacitance.
The linear graph is shown in Fig.5.16b.
5.3.1.2 Example Two
The fluid system shown in Fig.5.17 consists of a pressure
source in the form of a fluid reservoir, a ball valve on the

5 Fluid, Electrical, and Thermal Systems

278
Fig. 5.16a Nodes of distinct
pressures. b Linear graph of the
fluid system

b
C
P(t)

vent to
atmosphere

1 R1

Pump

Fluid
Reservoir

R1

p(t)

R2

R2

g
g

Fluid
Reservoir

Ball valve
Closed

Fig. 5.17 Fluid system with a fluid reservoir acting as a pressure


source, and fluid resistance, inertance, and capacitance

discharge of the reservoir, a fluid resistance R, a fluid inertance I, a fluid capacitance C, and second ball valve to drain
the system. A ball valve can be modeled as having negligible 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
this system, we will model open ball valves as having zero
fluid resistance.
Identify nodes of distinct pressure on the schematic,
Fig.5.18. The pressure at the discharge port of the fluid reservoir extends to the ball valve, when it is closed, and to the
fluid resistance R when the ball valve is open. As we are interested in the dynamic response when the ball valve is open,
lets take a look at the latter case. Pressure drops across the
fluid resistance, R, in the direction of the fluid flow. Hence,
pressure node 2 is immediately to the right of the fluid resistance. The fluid inertance I, must have pressure nodes at
either end. Node 2 is the pressure node for the left end of the
fluid inertance. Node 3 for the right end of the inertance is
in the fluid capacitance, which is connected to the fluid inertance by an ideal pipe.
Create the linear graph by drawing and identifying the
nodes. Next, draw the branches between them, per Fig.5.19.
This system is the fluid analog of the series RLC electric circuit of Sect.1.5.2.2. Most mechanical engineering
students find the fluid analogy with electric circuits helpful,
particularly as an aid for understanding electrical inductance,
the electrical analog of fluid inertance. This fluid system is
of second order, because the fluid inertance and the fluid
capacitance are independent energy storage elements. Con-

Ball valve
Open

Fluid
Reservoir

1 R2

3
I

Fig. 5.18 Fluid system with nodes of distinct values of the across variable, pressure

p(t)

Fig. 5.19 Linear graph of the fluid system shown in Figs. 5.17 and 5.18

sequently, its homogeneous, or natural, response can be oscillatory. The flow can slosh back and forth. Whether water
sloshes in a bucket depends on the relative amounts of power
dissipated in the fluid resistance and stored in the fluid inertance and capacitance.
Consider the initial condition of the ball valve on the reservoir as closed and fluid level in the capacitance below that
of the fluid reservoir, shown in Fig.5.17. If the ball valve
is opened quickly, the system experiences a step change in
pressure at node 1. How much flow is there in the system
at time, t = 0+, the instant after the valve is opened? Zero.
The mass of fluid in the fluid inertance must be accelerated
by the net pressure force between nodes 2 and 3. If the fluid
resistance is relatively low, then the flow through the inertance can build up to a torrent, storing a significant amount
of kinetic energy. On the other hand, if the resistance is relatively large, the volume flow rate in the inertance will not be

5.3 Linear Graphs of Fluid Systems

279

C
P(t)
vent to
atmosphere

R1

R1

Pump

R2

p(t)

R2

Fluid
Reservoir

Fig. 5.20 Fluid system schematic with a fluid power unit, modeled
as a pressure source, two fluid resistances, inertance, and capacitance

P(t)
vent to
atmosphere

1 R1 2

Pump

Fluid
Reservoir

C
3
R2

Fig. 5.21 Nodes of distinct values of the across variable pressure indicated on the schematic

as great, since the resistance restricts the flow through the


inertance, and less kinetic energy will be stored.
In the case of low resistance, when strong flow through
the inertance fills the capacitance up to the level of the reservoir, and a net back pressure is established to decelerate
the flow through the inertance, there is sufficient kinetic
energy stored in the inertance to maintain the flow into the
capacitance for some period of time. The fluid level in the
capacitance overshoots the reservoir level, which is the future equilibrium level, storing excess energy in the capacitor.
The flow then reverses back into the reservoir. The oscillation repeats, until the fluid resistance dissipates the energy
stored in the first cycle.

5.3.1.3 Example Three


The fluid system shown in the schematic of Fig.5.20 is driven by a pump, modeled as pressure source. The system has
two fluid resistances, R1 and R2, a fluid inertance, I, and a
fluid capacitance, C (fluid accumulator).
Identify locations of distinct values of the across variable, pressure, on the schematic and label them, Fig.5.21.
Properties are identified by the element parameters on the
schematic. If there is no property associated with a section
of pipe, then it is ideal, and there is no pressure change along
that section. The pipe segment from discharge port of the
pump to the fluid resistance, R1, is pressure node 1. Pressure
node 2, on the other side of fluid resistance R1, is also the

Fig. 5.22 Linear graph of the fluid system shown in Figs.5.20 and
5.21

node for the left end of the fluid inertance, I. The third pressure node is in the fluid capacitance, C. The fluid capacitance
is connected to the right end of the fluid inertance and the
top of fluid resistance, R2, by ideal pipes, so those locations
are also part of node 3. The pipe connected to the bottom of
resistance R2 has no properties and, thus, no pressure change
along it. The left end of the pipe discharges into the fluid
reservoir, which is at atmospheric or ground pressure, g.
The same pressure node extends along the pipe to the bottom
of resistance, R2.
Create the linear graph by first drawing and labeling
the nodes. Then, add the branches between their associated
nodes. Define a positive direction for the through variable,
volume flow rate. Sources are special, in that they raise the
value of the across variable, pressure, in the positive direction of the through variable. There is a decrease, or drop, in
the across variable, pressure, across resistances and energy
storage elements in the positive direction of the through variable, Fig.5.22.

5.3.1.4 Example Four


A common error drawing linear graphs from a fluid schematic occurs when dealing with a pipe, tube, or hose with
more than one energetic property. A linear graph is an energetic model. Elements in the linear graph represent the energetic properties of the system. If one component has two
energetic properties, then it will contribute two elements in
the linear graph. Consider a pipe that possesses both fluid
inertance and fluid resistance, per the schematic shown in
Fig.5.23.
The pipe extending from the discharge of the pump to
the fluid capacitance is called out as having two energetic
properties, fluid resistance and inertance. This is a common
circumstance. The properties of fluid resistance and fluid
inertance are both distributed along the entire length of the
pipe. They are two distinct properties in the same physical
component. The pressure drop along the pipe is due to both
the energetic properties, resistance and inertance. How does
one include both properties in a linear graph of the system?

5 Fluid, Electrical, and Thermal Systems

280

Long pipe with fluid


resistance and inertance

Node 2 divides pressure drop p13


between the pipes distributed
resistance R p and inertance I

P(t)
vent to
atmosphere

Pump

P(t)

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
pressure drop is divided, Fig.5.24a. The same pressure
drops across two elements in parallel, and the volume flow
rate, Q, is divided, Fig.5.24b. In this example, the two properties exist in one pipe. The energy dissipated in the fluid
resistance, and the energy stored in the fluid inertance stem
from the same volume flow rate, Q. The energetic properties of inertance and resistance are represented by elements
in series.
The total pressure drop along the pipe is due to both of
the energetic properties, resistance and inertance. The total
pressure drop must be divided by the addition of a node, i.e.,
node 2. Node 2 has no physical location, but it does have
physical meaning. Whether the resistance is assigned the
pressure drop from node 1 to node 2, or the inertance is assigned the pressure drop from node 2 to node 3, or vice versa,
is irrelevant. Both properties are actually distributed along
the length of the pipe. As long as both elements in the linear graph have the same volume flow rate, Q, through them,
and their pressure drops sum to the pressure drop along the
pipe, p13, the energetic properties are properly represented,
Fig.5.26.

5.3.1.5 Example Five


Linear graphs of fluid systems presented as more realistic
schematics are drawn by first reading the callouts on the
schematic to understand what energetic properties are in the
model. Pressure nodes are identified based on the properties
attributed to the components.
The fluid system shown in Fig.5.27 consists of a fluid
power unit modeled as a pressure source, two hydraulic
Fig. 5.24a Fluid resistance and
inertance in series. b Fluid resistance and inertance in parallel

a
1

Rp

R
g

Fluid
Reservoir

Fig. 5.23 Fluid system with two energetic properties, fluid resistance
and inertance, in one physical component, the long pipe

Pump

vent to
atmosphere

Fluid
Reservoir

Fig. 5.25 A fictitious pressure node in the pipe has no physical location. It divides the pressure drop, p13, into one pressure drop due to fluid
resistance and another pressure drop due to fluid inertance

hoses or lines, and a fluid accumulator. The properties of


the components are called out as follows. The hose from the
high-pressure pump or discharge port of the fluid power
unit to the fluid accumulator has fluid resistance R1. The
fluid accumulator has capacitance C. The hose from the fluid
accumulator to the low-pressure tank or return port has
fluid inertance I and resistance R2.
Pressure nodes are first assigned to locations on the schematic, which have distinct values of the across variable, pressure. These locations are the pump discharge port, the tank
return port on the fluid power unit, and the fluid accumulator. The hydraulic line from the accumulator to the return
port has the properties of inertance and resistance. Therefore,
a fictitious node is needed to divide the total pressure drop
between the two contributing properties, Fig.5.28.
The linear graph of this system is shown in Fig.5.29.

5.3.1.6 Example Six


Lets develop a model and draw its linear graph for a familiar
fluid system, a garden hose. Typical residential water pressure is 3040psi. We will model the residential water pipe
as a pressure source. Assume the following conditions. The
garden hose is attached to the sill tap (valve at the house),
which is closed. A nozzle on the end of the hose is also
closed. The hose is filled with water but remains depressurized, Fig.5.30.
If the sill tap valve is opened, with the nozzle still closed,
we would expect to hear some water flow into the hose,
which would swell slightly. The water, which flowed into
the hose, would be stored in a fluid capacitance created by
straining the hose circumferentially. If we next open the

I
2

R
I

5.3 Linear Graphs of Fluid Systems

281

Fig. 5.26 Equivalent linear


graphs for the system shown in
Figs.5.24 and 5.25

R1

p(t)

R2

p(t)

Fluid Power Unit Modeled as


Pressure Source p(t)

R1

R2

R1

Breather Cap (vent to atmosphere)


High Pressure Output
Hydraulic Line
Resistance R1

p(t)

Fluid Accumulator
Capacitance C

Low Pressure Return

R2

Fig. 5.29 Linear graph of the fluid system shown in Figs.5.27 and
5.28

Hydraulic Line
Inertance I and
Resistance R 2

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
inertance.

Sill Tap

p(t)

Hose
Inertance I
Resistance RH
Capacitance C

Nozzle
Resistance R N

Fluid Power Unit Modeled as


Pressure Source p(t)
Breather Cap (vent to atmosphere)

High Pressure Output

Hydraulic Line
Resistance R1

Low Pressure Return

Hydraulic Line
Inertance I and
Resistance R 2

Fluid Accumulator
Capacitance C

Fig. 5.28 Schematic annotated with pressure nodes. The fluid capacitance 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
distributed properties, inertance I and fluid resistance R2

nozzle, we may or may not be aware of a brief transient due


to the inertance of the hose. The water then flows in steadystate, with the volume flow rate a function of the sill tap pressure, hose resistance, and nozzle resistance.

Fig. 5.30 Garden hose attached to sill tap, which is connected to the
residential water system, modeled as a pressure source

The fluid resistance of the nozzle is easy to model as a


lumped parameter, because it is the nozzles only energetic
property. The hose has all three energetic fluid attributes of
inertance, resistance, and capacitance. All three properties
are distributed along the length of the hose. When we previously modeled a pipe with both fluid resistance and inertance, we found the properties to share the same volume flow
rate and, therefore, be in series. A fictitious node is needed to
divide the pressure difference between the two ends of a pipe
into the two pressure differences, one due to fluid resistance
and the other due to fluid inertance.
How is capacitance included in the model? The capacitance of the hose is distributed along the length of the hose,
due to pressure-induced circumferential strain which increases the cross-sectional area of the pipe. (There is also axial
strain that increases its length, which we will neglect.) We
must lump it into a single fluid capacitance element that is

5 Fluid, Electrical, and Thermal Systems

282

 Fig. 5.31a Pressure nodes for


lumped parameter model of a
garden hose. b A linear graph
corresponding to the nodes of
5.31a

Hose
Inertance I
Resistance RH
Sill Tap Capacitance C

p(t)

RH

quire the fluid capacitance node


to be the average pressure value
in the hose. b A linear graph corresponding to the nodes of 5.32a

b
1

RH1

I1

located between a node in the hose and atmospheric (ground)


pressure. We will begin by identifying where the hose capacitance cannot be located in the model. The energy storage
equation for a fluid capacitance is
1
2

EC = Cpxg2
where subscript x indicates the unknown location for the
lumped fluid capacitance. Capacitance is pressure-dependent
energy storage. The lumped capacitance cannot be located at
the sill tap, modeled as a pressure source. If we were to connect the capacitance to the pressure source, a step change
in pressure would instantly fill the capacitance, requiring an
infinite volume flow rate and an infinite fluid power flow.
There must be a fluid resistance or inertance between a pressure source and a fluid capacitance, to limit the volume flow
rate and power flow into the capacitance to finite levels.
Since there must be a fictitious node between the hose resistance and inertance, we will connect the capacitance there.
The schematic with nodes and a corresponding linear graph
is shown in Fig.5.31.
The linear graph of Fig.5.31b can be improved, at a cost
of increasing the number of elements. We can divide the

RH 2

I2

RN

p(t)

Nozzle
Resistance RN

RN

p(t)

Hose
Inertance I
Resistance RH
Capacitance C

Sill Tap

p(t)

Nozzle
Resistance R N

 Fig. 5.32a Pressure nodes re-

value of the hose resistance and inertance in half. We can


then place a resistance and inertance in series, both before
and after the node to which the capacitance is connected. The
advantage of the additional elements is that the capacitance
pressure node becomes the average pressure along the length
of the hose, when flow is steady and the inertance, therefore,
does not affect pressure, Fig.5.32.

5.4Calculating Fluid Element Parameters


from Fluid Properties and Geometry
5.4.1 Fluid Resistance
Similar to friction in mechanical systems, it is generally
more difficult to calculate fluid resistance than either fluid
inertance or fluid capacitance, excluding capacitance due to
fluid compression. The difficulty stems from having to define the non-linear behavior of real fluids, such that we can
calculate values for a resistance independent of fluid velocity
in the linear form, p12 = RQR.
In the ideal case of steady flow in straight and perfectly
smooth piping, the Reynolds number, Re = vD / , where v
is the representative velocity of the flow and D is the diameter

5.4 Calculating Fluid Element Parameters from Fluid Properties and Geometry
Table 5.1 American Petroleum Institute (API) Engine Oil
Classifications 2004. (Source
SAE J300). W (from Winter)
indicates the low temperature
viscosity

Low-shear rate
High-shear rate
SAE
Viscosity Minimum kinematic viscosity Maximum kinematic viscosity Maximum viscosity
rating
mm 2
mm 2
at 100C
atv 100C
1cP = 1mPa sec at 150C
cSt =
cSt =
sec
sec
0W
5W
10W
15W
25W
20
30
40

3.8
3.8
4.1
5.6
5.6
9.3
5.6
12.5

<9.3
<12.5
<16.3

40

12.5

<16.3

50
80

16.3
21.9

<21.9
<26.1

of the pipe, tube, or hose, is used to determine if the flow is


laminar or turbulent. By definition, the transient portion of
the dynamic response of a fluid system has changing flow
rates and hence changing Reynolds numbers. Consequently,
a representative Reynolds number cannot be estimated, without preliminary dynamic modeling to provide approximate
flow rates. Fluid flow is also sensitive to relatively smallscale, geometric details of the flow path and surface condition (roughness) of the components.
The typical case of a machines hydraulic system is characterized by the hydraulic power unit (pump)s turbulence,
sustained by the fluids passage through a variety of tees,
valves, fittings, and actuators. Turbulent flow can be considered independent of Reynolds number, since it is slug
flow, Fig. 5.2a. However, the Reynolds number of the flow is
needed to estimate the friction factor, f, which is used to calculate the Resistance Coefficient K and its corresponding
Equivalent Length of a straight pipe, which would yield
the same pressure drop. The Colebrook equationis used to
calculate the friction factor, f , for turbulent flow with Reynolds greater than 4,000.

1


2.51

= 2 log D +
3.7 Re f
f

283

where roughness

(5.30)

Published design values for the roughness of steel piping, tubing, and hydraulic hose vary considerably. Cited values for
new steel piping or tubing have ranged from = 0.0006 in.
to an order of magnitude lower at = 0.00006 in. The span
of published values for new hydraulic hose is greater, from
= 0.001 in. to = 0.0003 in. This variation reflects the fact
that hydraulic hose is a manufactured product, whereas steel
tubing is a processed material. The roughness of the interior

2.6
6.9
2.9
(0W-40, 5W-40, 10W-40)
3.7
(15W-40, 20W-40, 25W-40, 40)
3.7
3.7

surface of the hose depends on both the surface material, an


elastomer (synthetic rubber), and design of the tubes fiber
reinforcement. The fiber reinforcement can impart waviness
to a surface which varies with internal pressure. One rarely
sees straight hydraulic hose. Hose is typically used to allow
flexure at pivots, where straightness of the hose depends on
the position of the machine. Hose is also employed for convenience in manufacturing. In the latter case, the hydraulic
hose is run like cable, where it will fit. Consequently, it may
have a circuitous run.
Although we cannot place confidence in fluid resistance
values calculated a priori, we have no choice, in the absence
of test data. A reasonable strategy is to estimate fluid resistance at the presumed middle of the operating range of flow
rates. One estimate is for a best-case scenario. Another estimate assumes unfavorable limits for the viscosity and geometry, where unfavorable depends on the function of the
system being modeled.
The ISO oil standards identify oil grade by viscosity in
centistokes at 40C. Hydraulic oil viscosities in the range
of ISO grades 32100 are typical. The viscosity ranges for
engine oils (also called motor oil, if the lubricating oil used
in internal combustion engines) is expressed as weight in
Society of Automotive Engineers (SAE) standards. Multiple 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.
Refer to Table5.1.

5.4.2 Fluid Inertance


As is the case with mechanical systems, we typically have
greater confidence in the parameter of the kinetic energy

5 Fluid, Electrical, and Thermal Systems

284

storage than in the systems other parameter values. The only


uncertainty in calculation of fluid inertance, I
I=

L
A

is the density of the hydraulic oil. Although there can be


significant variation in viscosity among hydraulic oils, their
density is much less variable. Lubricating oils range in density from approximately 850 to 950kg/m3. Hydraulic oils
range approximately from 860 to 900kg/m3. Emulsions are
denser than mineral oils, due to their water content. The
above densities are uncertain due to the unknown amount of
air entrained in hydraulic oils. The amount of entrained air
can reach 9% by volume at atmospheric pressure.

5.4.3Fluid Capacitance (Hydraulic


Accumulators)
Fluid capacitance C is easiest to calculate using the integral
form of the elemental equation
QC = C

dp4 g
dt

QC dt = Cdp4 g

Vol = C p4 g

p2
p1

C=

C=

Vol
p4 g

t2

p2

t1

p1

QC dt = C dp4 g

Vol
p4 g2 p4 g1
(5.31)

The bulk modulus, , of hydraulic oil lies in the range of


250,000 to 300,000psi. This is attributed to the unknown
amount of entrained air and its dependence on temperature
and pressure. The far greater bulk modulus of hydraulic oils
makes it preferable to use nitrogen as a gas spring in hydraulic accumulators.
The most commonly used type in machine design is the
nitrogen gas-charged bladder-type fluid accumulator. The
pressure of nitrogen gas is a percentage of the design minimum hydraulic fluid pressure, pmin, and the intended function
of the accumulator. The engineer would choose the highest
charge pressure, typically 80% of pmin, if the accumulator
is intended to store energy, in order to allow the use of a
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 system, to absorb a shock loading after the sudden stoppage of
a large actuator. A pressure between those limits is selected,
if the fluid capacitance is intended to smooth out pressure

fluctuations or pulsations in the system. An electric capacitor


is used in the same way to smooth the voltage produced by a
diode bridge in a DC power supply.
Gas-charged bladder-type accumulators, Fig.5.11, resemble gas cylinders. They are described by two different
volumes. The larger volume of the two is the nominal volume of the entire interior volume of the pressure vessel. The
more important and smaller volume is the effective volume. It is the maximum volume of hydraulic fluid, which
can be expelled by gas-charged bladder, when fluid pressure
is varied from pmax to pmin. The nominal and effective volumes are essentially equal for small accumulators, but differ
by up to 10% for larger accumulators, because the cylinder
volume is increased by increasing the length rather than diameter. This limits the expansion of the gas bladder.
Gas-charged diaphragm-type accumulators resemble kettles. The elastomeric diaphragm is attached to the interior
circumference midway up the side. Gas-charged diaphragmtype accumulators have equal nominal and effective volumes.
Gas-charged piston-type accumulators have advantage
over gas-charged bladder or diaphragm types of greater discharge when used to store energy because their charge pressure is higher: 100psi below pmin, rather than 80% of pmin. If
pmin were 2,000psi, the difference in charge pressure would
be 300psi between the piston-type accumulators and the
others. It is possible to find gas-charged piston-type accumulators in more combinations of diameter and length than
are available for gas-charged bladder or diaphragm types.
They do have frictional resistance due to the seals between
the piston and cylinder, but not as much as the spring-loaded
piston-type accumulator, because the pressure drop across
the piston is not as great.
The spring-loaded piston-type fluid accumulator is the
least common type of hydraulic accumulator. The frictional
resistance of the piston-cylinder seals which, as a rule of
thumb is 10% of the rated net pressure force, is the same as
that of a hydraulic piston-cylinder actuator.

5.5Fluid Power System Hardware


and Symbols
Most fluid power schematic symbols are self-explanatory.
There are three worth noting. Of these, two are self-explanatory, when their functions are understood. The third is a
puzzle, until its corresponding hardware is introduced.

5.5.1 Metering or Flow Control Valves


Metering or flow-control valves are either fixed fluid resistances, per Fig.5.33a, or manually adjustable fluid resistanc-

5.5 Fluid Power System Hardware and Symbols

285

a
1

Fig. 5.33a Fixed resistance metering or flow control valve, b Adjustable metering valve

b
1

2
3

Flow lifts ball from valve seat

Flow

c
No Flow

2
3

Flow seats ball closing valve

Fig. 5.34 Check valve hydraulic schematic symbol

es, per Fig.5.33b. The base symbol is a flow restricting to


indicate a fluid resistance. The diagonal arrow through the
adjustable valve follows the same convention used to indicate adjustable electrical components.

5.5.2 Check Valves


Check valves are used to limit backflow in fluid systems.
They are the fluid analog of an electrical diode, since they
permit flow in only one direction. There is also a pressure
drop in the flow direction through a check valve, analogous
to the voltage drop with flow of current through a diode. The
schematic symbol for a diode is a circle in an angle, which is
intended to imply a ball-type check valve, in which backflow
pushes a ball against the valve seat, closing the valve, per
Fig.5.34. Ball-type and poppet-type check valves are used
in hydraulic systems. A poppet valve has a tapered cylinder
or cone, rather than a ball, held against the valve seat by a
spring. In low-pressure fluid systems, most check valves are
disk type, in which a disk is held open against a spring by
the forward pressure of the flow. Flap valves are also used as
check valves in low-pressure systems.

5.5.3 Multi Position Shuttle or Spool Valves


Shuttle and spool valves are important in machine design, due
to their use to control the direction of fluid power flow. These
valves fall under the category of cartridge valves, since the
overall shape of the mechanism fits into a cylindrical cavity in the valve body or manifold. They are multi position
valves, whose schematic symbol is a puzzle to the uninitiated,

Fig. 5.35 Three-position spool valve. The schematic symbol for each
position is shown to the left. A cross-section through the valve is shown
on the right. a Flow from port one to port two. b All ports blocked.
cFlow from port two to port three. In practice, all valve positions are
shown on schematic as in Fig. 5.36

because it shows the function of each position of the valve,


Fig.5.35.
The name shuttle valve refers to components moving,
or shuttling back and forth, within the valve body. This action serves to block or reveal ports and, thus, control the
flow. A poppet-type valve sports a hollow, cylindrically
shaped poppet, which is closed at one end. A taper at the
closed end of the cylinder, along with a spring within the
cylinder to push the taper against the valve seat, makes it
a poppet valve. The diameter of the cylinder adjusts, to either block fluid or allow flow when aligned with ports in the
valve body, as described.
Spool valves are so named, because the moving component, which controls the flow, resembles a spool for thread.
The spool is a precision-machined, cylindrical part of varying diameters. When a small diameter portion of the valve is
aligned with pressure and discharge ports in the valve body,
there is flow through the valve. When a full diameter portion aligns with the flow and discharge ports, the flow is
blocked. Multiple positions of a shuttle or spool valve allow
one valve to control the movement of a hydraulic piston/
cylinder.
When spool valves are part of an automatic system, they
are called servo valves. The prefix, servo, means slave.
The most precisely manufactured servo valves are known as
linear, or proportional, valves. They are designed and
manufactured, so the pressure drop through the valve is linearly related to the position of the spool. Linear valves are

5 Fluid, Electrical, and Thermal Systems

286

Fig. 5.36 Schematic symbol for the three-position spool valve shown
in Fig.5.36. The schematic symbol shows the flow control for each
position of the spool

used in closed-loop feedback controls of hydraulically powered machines.

5.6 Electrical Systems


We will introduce elements of electrical systems, drawing
on analogies to fluid systems. We will also investigate the
phenomenon of magnetic energy storage, which results in
inductance.
The electric field and the interaction of electric charges
are both familiar and mysterious. A familiar aspect is the
force exerted between socks, when they are removed from
the dryer. It is a substantial force, far stronger than gravity,
or we wouldnt have to peel socks apart, as they would easily
separate, when we pick up one of the pair. The collection of
electric charge, the resulting voltage, and electric field are
created by polymeric, synthetic fibers tumbling against one
another in the dryer. The voltage is created by the mechanical separation of charges.
A mysterious aspect of the electric field is that it is apparently an aspect of space itself, and exists at a random nearzero level, when no charges are in the vicinity. As bizarre as
it sounds, if the electric field were created by charge, then
we would know that the field would be exactly zero, if no
charge was in the vicinity. The Heisenberg uncertainty principle prevents us from knowing the electric fields strength.
If the strength of the electric field must be unknown, it cannot be zero. Strange, but apparently true.
A circuit which is powered (or excited) by a constant voltage or current is known as a DC (direct current), even though
the voltages and currents within the circuit may oscillate during the transient period of the response to a step change in
applied voltage or current. The steady-state response of a DC
circuit is a constant voltage or current.
The steady-state response of a circuit powered by a sinusoidal voltage is sinusoidal. We refer to a circuit as an AC
(alternating current) circuit. It is possible to power some circuits with either a constant voltage or a sinusoidal voltage.
However, in practice, AC circuits are designed to be powered by sinusoidally varying voltage sources. (DC circuits

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.
We must defer further discussion of AC circuits, until we
have developed the theory needed for frequency response,
which is the steady-state response of a system to a sinusoidal
input. We shall see that the response of a circuit to a sinusoidal voltage source can be extended to any voltage with
periodically varying (repeating) voltage, by means of a Fouriers series approximation. Consequently, we will broaden
our definition of DC circuits to be all circuits except those
excited by a periodic voltage of any shape (or waveform).

5.6.1Analogies Between Fluid and Electrical


Systems
Fluid systems are mechanical systems, since they have
mass and are described by Newtons laws. The physical
laws, which describe electrical systems, were discovered
after Newtons time. Although electrical systems involve
forces, the mass of the charge carrier is generally neglected.
However, there are strong mathematical analogies between
electrical and fluid systems. Many mechanical engineering
students find these analogies very helpful in understanding
electrical systems.
The mathematical analogies are apparent in a comparison
of the elemental and energy equations of fluid and electrical
systems, Table5.2. Both fluid and electrical systems contain
a quantity which flows through elements and is conserved.
The flows of the quantity are driven by a potential drop. A
pressure drop between two ends of a pipe drives fluid flow
through the pipe. Likewise, a voltage drop between two
ends of an electrical resistor drives current flow through the
resistor. The energy dissipation element in both systems is
called a resistor. Whether the element is a fluid or an electrical resistor is apparent with the identification of either fluid
or electrical variables in the equation. Similarly, the element
that stores energy as a function of the across variable, either pressure or voltage, is called a capacitor in both types
of systems.

5.6.2 Summary of Electromagnetic Phenomena


The following summarizes the fundamental electromagnetic
phenomena and definitions we need for the subsequent discussion:
Electric charge is conserved.
Electric charges of opposite polarities exist with the electron and proton carrying unit elemental negative and

5.6 Electrical Systems

287

Table 5.2 Comparison of fluid and electrical system equations


Fluid system
Electrical system
P
= v12i
P
= p12Q
Power
Energy dissipation
Across-variable
Energy storage

Fluid resistance
p12 = RQR

Electrical resistance
v12 = RiR

Fluid capacitance

Electrical capacitance

QC = C

dp1g

dv12
dt
1 2
EC = Cv12
2

iC = C

dt

1
2

EC = Cp12g
Through-variable
Energy storage

Fluid inertance
p12 = I

EI =

Electrical inductance

dQI
dt

v12 = L

1 2
IQI
2

EL =

diL
dt

1 2
LiL
2

positive charges, respectively. The magnitude of the elemental charge was determined by Millikan, who quantified the electric force acting on miniscule oil droplets
falling between charged plates.
The unit of electric charge is the coulomb, C, which
is defined in terms of the flow of an electric current in
amperes as


ampere

coulomb
second

(5.32)

1 coulomb = 6.2 1018 elementary or electronic charges.


A coulomb is an enormous amount of charge. In practice,
it is difficult to isolate such a large charge.
The unit difference in electrical potential or electromotive
force is the volt, V, defined the base SI units as


volt

kg m 2
sec3 ampere

(5.33)

Rewritten in derived SI units,


volt =



kg m 2
sec3 ampere

The ampere is defined by the magnetic force acting


between two straight conductors, specifically, the ampere
is that constant current which, if maintained in two
straight parallel conductors of infinite length, of negligible circular cross-section, and placed one meter apart in
vacuum, would produce between these conductors a force
equal to 2107N per meter of length.
Modeling two collections of charge, q1 coulombs and
q2 coulombs, as located at points, the force exerted on a
static or moving charge q1 by a static point charge q2 is
calculated by the inverse square law:

kg m

1
volt =
m
2
sec ampere
sec

Felec =

1 q1q2
r12 newtons
4 r12 2

(5.36)

where is the electrical permittivity, a property of the material between the charges. The factor, 4, was included to
aid calculations, since an inverse square law force incorporates spherical geometry. The permittivity of free space, a
vacuum, is


0 = 8.85 10 12

coulomb 2
N m2

(5.37)

N m2
coulomb 2

(5.38)

which yields


1
4 0

= 9 109

A moving electric charge, q, creates a magnetic field, B,


which is proportional to the magnitude of the charge, q,
and the velocity of the charge, v.
A static electric charge, q, is unaffected by a magnetic
field, B.
A magnetic field, B, exerts a force on a moving charge
perpendicular to both the velocity, v, and the uniform
magnetic field, B, or, in other words, normal to the plane
defined by v and B. The magnetic force is calculated as a
cross product


Fmag = q v B

(5.39)

1
volt = N m
coulomb

(5.34)

Note that this implies that no magnetic force acts on a charge,


q (or current, i), moving parallel to the magnetic field, B,
since

volt coulomb = joule

(5.35)

Moving a coulomb of charge against the electrical potential


of one volt requires 1J of energy. Likewise, one coulomb of
charge at one volt has the potential energy of 1J, hence the
term, electrical potential.

Fmag = q v B sin

(5.40)

where is the angle between velocity, v, of charge, q, and the


magnetic field, B.

5 Fluid, Electrical, and Thermal Systems

288

5.6.3 Electrical Units


Historically, electrical and magnetic phenomena were discovered by their mechanical effects. Gauss (17771855)
championed the cgs (centimeter-gram-second) system of
measurement, in which electrical variables were derived
from mechanical variables. This system was also used by
Maxwell (18311879) and Thomson (Lord Kelvin) (1824
1907). It was very convenient mathematically, but not practically. Unfortunately, the derived electrical units depended on
whether the electrostatic or the magnetic phenomenon was
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
electrical units are SI units. The SI unit of current, the ampere, was adopted in 1948. The SI unit of electric charge, the
coulomb, is derived from the ampere. One volt will drive one
coulomb (6.241018) of charge carriers, such as electrons,
through a resistance of one ohm in one second. One coulomb
of charge carriers, moving across a surface or through a node
in one second, is one ampere of current.
The sign convention for electrical systems was established by Ben Franklin, long before the electron was discovered. Franklin had to guess which way the charge carriers,
electrons, were flowing; from positive to negative or from
negative to positive. He guessed wrong. We still follow his
convention of current flowing from positive to negative,
even though we now know that electronsthe most common charge-carrierare negatively charged. Consequently,
when current is composed of electrons, as is almost always
the case, the flow of charged particles is from negative to
positive, which is the opposite direction of current. The type
of charged particle, and, consequently, the direction of the
flow of charged particles, is inconsequential in the macroscopic model of electrical systems. It is of great consequence
on the microscopic scale. Metals have low electrical resistance to the flow of electrons but high resistance to the flow
of protons or ions. Aqueous solutions and molten salts, on
the other hand, can transmit either type of current.
The SI definitions of electrical properties are logical but
result in awkward magnitudes. For example, the property of
capacitance is expressed in farads with the symbol, F.


farad

volt
volt
farad
coulomb
ampere second

(5.41)

A one farad capacitor can store one coulomb of charge at 1V.


Logical, but a one farad capacitor is an enormous capacitor.
The common unit for capacitance used in electronic circuits
is a picofarad, pF, where pico is 1012. Electrical inductance
is the voltage developed across an electrical conductor, in
response to the rate of change of current of one ampere per
second. The unit of inductance is the henry, H, named for

Joseph Henry (17971878), who discovered self-inductance


and invented the electric relay. The plural of henry is either
henrys or henries.


henry

volt
volt second
=
ampere
ampere
second

(5.42)

An inductance of one henry is huge. The values of electrical


inductors used in electronic circuits are typically expressed
in microhenries.

5.7 Electrical System Elements


There are two energy storage elements in electrical systems.
Electrical capacitors store energy in an electric field created
by opposite electric charge, and collected on parallel conductive surfaces. Inductors store energy in the magnetic field,
established by moving electric charges. The energy dissipation mechanism in electric systems is electrical resistance.

5.7.1 Electrical Resistance


Electrical conductivity is the ease with which electric charge
moves through a material. Conductivity and its inverse,
electrical resistivity, are material properties. Most materials
have negligible electrical conductivity. Of solids, the materials with the highest electrical conductivity at normal temperature are metals, which form crystal lattices, and share
conduction electrons. Silver is the most conductive, marginally more conductive than copper, followed by gold and
aluminum. Silver and gold are used to plate contacts. Silver
acts to resist damage due to arcing, and gold acts to prevent
oxidation. Copper and aluminum are used as conductors.
Semiconductors are metals with orders of magnitude less
conductivity than silver and copper, hence their name. Silicon and germanium are the two, which are most widely used
in electronics manufacturing. Carbon is conductive. The first
electric lights were arc lights with carbon electrodes. Edisons first light bulb used a carbon filament. Carbon is used
in its graphite form for motor brushes, which are the electrical contact between the stationary frame of the motor, and
the rotating commutator contact on the motors shaft.
The thermal motion of mobile conduction electrons in
metals is akin to the random motion of molecules in a gas.
A voltage on a conductor creates an electric field, which imposes a force on the conduction electrons, thereby accelerating them and bending their random motions in the direction
of the electric field. The acceleration is modest, compared to
the great speed of the conduction electrons random motion.
However, the average velocity of the electrons in random

5.7 Electrical System Elements

289

60 ft of 24 AWG
solid copper wire

1.5 VDC

motion is zero, since they are moving in all directions. The


acceleration imposed by the electric force is sufficient to
create a non-zero average velocity called the drift velocity. The drift velocity is the velocity of the current. Collision between the conduction electrons and the copper lattice
of the conductor transfers energy from the electrons to the
metal lattice. This energy transfer is in equilibrium, when the
electrons are in thermal motion. However, when an external electric field accelerates the conduction electrons, there
is net energy transfer to the lattice during collisions, which
results in heating of the metal conductor. This is the mechanism of electrical resistance. Defects in the crystal lattice increase the frequency of collisions. Electrical grade copper is
99.99% pure. High purity is needed to limit disruption of the
copper crystal lattice by contaminants.
Electrical resistance is the inverse of conductivity. Increasing electrical resistance decreases the current produced by a
voltage. Decreasing the cross-sectional area of a conductor
increases its electrical resistivity. Increasing the length of a
conductor increases its resistivity. This is directly analogous
to how the fluid resistance of a hose changes with its geometry. If the cross-section of a hose is reduced, there will be
less flow for a given pressure drop. If the length is increased,
there will also be less flow.
Two symbols are used for electrical resistance, per
Fig.5.37. The US symbol is a zigzag, which resembles the
coil of toaster or a strain gauge. The European symbol is a
rectangle.

5.7.1.1 Resistive circuit calculation


An electric circuit shown in the schematic, Fig.5.38. Two
1.5 VDC AAA batteries in series are connected to a 2 resistor by 60 feet of 24 AWG (American wire gauge) solid
copper wire. Calculate the current through the circuit and the
voltage drop across the 2 resistor.
In this calculation, the schematic may not represent our
model. Normally, conductors in schematics are considered
ideal and without resistance. In this circuit, resistance of the
discrete resistor is relatively small and may be on the order
of the wire. We need to calculate the resistance of 60 feet of
the 24 gauge solid copper wire in series with the 2 resistor.
The electrical material property analogous to magnetic
permeability is electrical conductivity. Electrical conductiv-

1.5 VDC

Fig. 5.37a US resistor symbol, b European resistor symbol

Fig. 5.38 Resistive electric circuit

ity is not generally reported; its inverse, electrical resistivity


is. Copper has resistivity of:

copper =

1.7 10 8
m

The unit of electrical conductance used to be mhos (ohms


spelled backwards) until SI was introduced. It is now siemens. Inverting the resistance of copper yields a conductivity of:

copper =

5.9 107
m

1
7
= 5.9 10 siemens
m

The resistance of a given piece of wire is


Rwire =

Length

copper Area

AWG (American Wire Gauge) 24, single-stranded (i.e.,


solid) wire has a diameter of 0.020in. The cross-sectional
area in square meters is
Area =

1m
0.020 in
= 2.03 10 7 m 2

4
39.37 in

[Aside: The diameter of gauge of wire in AWG is not a single


number. The diameter of a gauge varies with the number of
strands in the wire.] Multiple-stranded wire has two advantages in machine design. The primary reason for the use of multiple-stranded wire is its long fatigue life. Even if a solid copper
conductor is not intended to flex during the operation of a machine, a conductor is subject to fatigue, if it is free to vibrate
when excited by a machines motion. At best, a fatigue failure
leads to arcing and then an open circuit. At worst, as the crack
develops, the increased resistance increases the resistive heating, resulting in a fire. The secondary reason is that, although
we model electrical current as conducted uniformly over a

5 Fluid, Electrical, and Thermal Systems

290

Equivalent resistance
of 60 ft of 24 AWG
solid copper wire

1.2

i R1
i

3V

Energetic Equations
Continuity: Node 1 i = iR1 Node 2 iR1 = iR2
Compatibility: v13 = v12 + v23
Elements: v12 = R1iR1 v23 = R2 iR2
Energy: No energy storage elements.
Reduction Input: v13 = 3VDC, Output: v23

i R2

v13 = 3 VDC = v12 + v23

Fig. 5.39 Resistive electric circuit with the equivalent resistance of


60ft of 24 AWG solid copper wire

cross-section of a conductor, the current density is higher at


the surface of a conductor. Multiple-stranded wire has greater
specific surface area, which is the ratio of the surface area to
the volume. Hence, more current can be conducted by less
copper. For example, a given length of AWG 24 wire comprising 41 strands for AWG 40 wire weighs 5% less than the
same length of single-stranded AWG 24 wire and has 5.7%
less resistance.
The length in meters is
1m
Length = 60 ft
= 18.3 m
3.28ft

Length
18.3 m
=
7
Area 5.9 10 siemens
7
2

2.03 10 m
m

18.3 m
1

5.9 107

2.03 10 7 m 2

Rwire =

v
3 VDC = R1 23 + v23
R2
R1 + R 2
3 VDC =
v23
R2

R1
+ 1 v23
3 VDC =
R2
R2
v23 =
3 VDC
R1 + R 2

The fraction of voltage dropping across an individual resistor


in a series of resistors is the ratio of that resistance to the total
resistance of the series. You will find this result intuitive, if
you make the analogy between electrical resistors and fluid
resistances, e.g., steady flow through two garden hoses in
series.
Substituting in values for resistances, R1 and R2, yields the
voltage drop, v23, across the 2 resistor,
R2

2
3 VDC =
3 VDC = 1.7 VDC
v23 =

1.5 + 2
R1 + R2

Yielding an electrical resistance of

Rwire =

3 VDC = R1iR1 + v23

3 VDC = R1iR 2 + v23

Rwire =

Length
= 1.5
Area

The electrical resistance of the wire is the same order of


magnitude as that of the resistor. Our model for the circuit is
shown in Fig.5.39:
Two resistors in series are called a voltage divider because the voltage across the pair drops across an individual
resistor, in proportion to its contribution to the sum of the
resistances. We can now calculate the voltage drop across the
2 resistor.

5.7.2Electrical Capacitance: Energy Stored


in an Electric Field
Fluid systems can store energy as a function of pressure,
by straining the container (the fluid accumulator), straining
(compressing) the fluid, or by pushing the fluid against gravity. In all the three cases, we visualize pressure pushing fluid
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 electric field of the capacitor is analogous to the energy stored
in the strain of fluid in a fluid capacitor. Work is done to
create the field, by moving like charged charge carriers into
proximity, just as work must be done to compress a fluid, by
moving the fluid molecules closer together.
Capacitance exists wherever charge exists. Charge affects
the electric field, E, and is affected by the electric field created by other charges, whether the charge is stationary, or
moving as a current. In order for current to flow into an electrical capacitor, the charge flowing in must be pushed in, to
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


Fig. 5.40 A parallel plate
electrical capacitor. Note that the
electric current flows through the
capacitor even though no charge
passes between the plates

291

iC

iC

Current is shown flowing through the capacitor. In reality,


no charge-carrier crosses the dielectric (insulator) between
the parallel plates. The electrical current through the capacitor consists of oppositely charged charge carriers flowing in
opposite directions. The two opposite signs cancel for the
negative charge carriers (electrons) moving opposite the current direction, resulting in a positive current flowing in the
positive direction.
Two styles of capacitors are in common use. Figure5.41a
is a schematic of a parallel plate capacitor. Old radios were
tuned by turning a knob that swung a stack of parallel aluminum plates on a spindle between a fixed stack, thereby varying the total area between the plates and the capacitance of
the variable capacitor, Fig.5.41b. Non-polarized capacitors, i.e., capacitors which do not have positive and negative
terminals marked on them, used in electronics today have
similar designs. These capacitors have conductors separated
by an insulator, a dielectric. One contemporary design uses
a thin film of plastic, which has been aluminized to make its
surfaces conductive. The aluminized surfaces are the plates,
and the film is the dielectric. A second film for insulation is
added. The film is then folded, or rolled up, and encapsulated. The capacitor has a large surface area with a small gap
between the plates.
A second type of capacitor in use is an electrolytic capacitor. The symbol for an electrolytic capacitor has an arc,
in place of one of the parallel lines of the plate style capacitor. The negative terminal of the capacitor is the side with the
arc, per Fig.5.41c. Electrolytic capacitors have a single aluminum foil plate. A piece of paper is placed on top of the foil.
The stack is rolled up, placed in a small cylindrical can, filled
with an electrolytic solution, and then sealed. The electrolyte
acts as the second plate of the capacitor. Electrolytic capacitors are used for large capacitances. Electrolytic capacitors
are polarized, because of the specific ions in the electrolyte.
Heed the polarity markings on the electrolytic capacitors.
If you are lucky, then you only have a non-working circuit
and leaked electrolyte to cleanup. However, capacitors are
known to have exploded due to reversed polarity.
Analogies between fluid and electrical systems are not
perfect. Fluid capacitance and electrical capacitance have two

Parallel Plate
Non-Polarized

Adjustable
Non-Polarized

Electrolytic
Polarized

Fig. 5.41 Capacitor symbols. aThe parallel plate symbol is used for
non-polarized capacitors. b An arrow through an electrical symbol indicates the device is adjustable. c An electrolytic capacitor. The curved
line is the negative terminal, the cathode

non-analogous aspects. First, fluid capacitors must be referenced to the environments ground pressure, usually atmospheric pressure. There is an absolute measure of zero pressure, a perfect vacuum. It is impossible to draw fluid from a
perfect vacuum. Conversely, there is no absolute measure of
voltage. Voltage is always relative. Measurements of voltage
may be made relative to chassis ground, which is the ground
used by a power supply in an electrical or electronic device.
Voltage measurement may also be made relative to earth
ground, which is usually the ground of the buildings electrical wiring. The reference used to calculate the energy stored
in the capacitance is the difference in voltage between the two
plates.
The second difference between fluid and electrical capacitance concerns their through variables. The through
variable of a fluid capacitance, volume flow rate, does not
flow through the capacitance. The fluid is either temporarily
stored in the capacitance, as in the case of a fluid accumulator, or the hose, tube, or pipe is strained creating a temporary increase in volume. Electrical capacitors have a flow of
positive charge in the positive direction to the positive plate,
and negative charge in the negative direction to the negative
plate. The negative signs of the charge and the direction of
the charge flow cancel, leaving an apparently positive current flowing across the gap between the plates. This current
is displacement current. Maxwell conducted experiments
which demonstrated that displacement current has all of the
magnetic effects associated with conventional current.

5.7.3Electrical Inductance: Energy Stored


in a Magnetic field
Electrical inductance is more difficult to understand than
electrical capacitance for three reasons. First, we can sense
the presence of electric fields, because we carry electric
charge on our bodies and clothing. Everyone has felt the
static electricity created by separating clothing that was

5 Fluid, Electrical, and Thermal Systems

292

iL

v1

Torrodial ferrite core


Relative permeability r
Core length l c
Cross-sectional area A

v2

iL

N turns of wire in the coil


Fig. 5.42 A Toroidal Inductor

clinging together. (Static electricity is created by relative


motion of dielectric surfaces in contact, which can result in
charge transfer between the surfaces. The mechanical work
done to separate the surfaces produces a voltage.) We are unable to directly sense magnetic fields. Consequently, we are
oblivious to them and unaware of their many effects.
Secondly, electric fields are created by the collection of
like charged charge carriers, whereas magnetic fields are created by the motion of charge carriers. Static phenomena are
typically easier to understand than dynamic phenomena.
Thirdly, the electric forces between static charge carriers
act in a straight line on the axis between the centers of the
charge carriers, and are easy to visualize. Two vector crossproducts are needed to describe the magnetic forces acting between two interacting, moving charge carriers. Consequently,
magnetic forces are difficult to visualize.
The magnetic energy storage property of electrical circuit elements is expressed as inductance L with the unit
of henry, H, Fig. 5.43. The parameter symbol L is in honor
of Heinrich Lenz, credited with Lenzs law. Henry was an
American scientist active in the early 1800s.
An induced current is created by a time varying, or
moving, magnetic field. Lenzs law states that an induced
current creates a magnetic field in opposition to the magnetic
field, which created the current. This is a statement of energy
conservation. If the induced current fed back and increased
the magnetic field, the process would create a magnetic field
of infinite energy.
Inductance L is a measure of how large a magnetic flux a
given current i can create. Magnetic flux is the component
of magnetic field B, normal to an area integrated over that
area. The term, flux, is used to evoke a flow analogy, since
the magnetic field lines form closed curves. The intensity
of the magnetic field increases with current, the number of
turns of wire in the coil, and the relative permeability of the
material in the coil. The turns of wire, which form the coil,
are said to be linked to the magnetic flux they create,
since the magnetic flux passes through the turns of the coil.
Figure5.42.
Magnetic permeability is a material property. Quotes
are used, because free space, the physics term for a complete

vacuum, has magnetic permeability. Ferromagnetic materials, iron, nickel, cobalt, and their alloys, increase the strength
of a magnetic field, due to the alignment of the spin axes
of their atoms with the applied magnetic field. Ferrite is a
ferromagnetic ceramic. Its ceramic aspect makes its electrical conductivity very low, which is desirable to reduce the
eddy currents induced by time varying, or moving, magnetic fields. Eddy currents heat the conductors, in which they
are formed. Temperature of the motor conductors is generally the factor, which limits the current and, hence, power an
electric motor can produce.
This coupling is called the flux linkage, and is given
symbol (Greek for l as in linkage). The flux linkage between a magnetic flux and a coil is proportional to the
number of turns of wire in the coil, N, the flux flows through
or threads (as in threading the eye of a needle) or links
(as in links of a chain).


= N

(5.43)

If the relationship between the magnetic flux threading a


coil and current i through a coil is linear, then there is a linear
relationship between the flux linkage and current i, which
defines the parameter inductance L.


= Li

(5.44)

Differentiating this expression with respect to time yields:


d
di
=L
dt
dt
A voltage v12 develops across an electrical inductor in response to a time-varying current i, or, in truth, due to the
time-varying flux linkage, . The current through the coil resists change because energy must flow into to coils magnetic field as the current increases and, similarly, flow out of the
coils magnetic field as the current decreases. The voltage
acts to drive the change in the current through the coil as the
current resists change. The voltage across a coil is analogous
to the force acting to accelerate a mass. The force drives the
change in the kinetic energy in the mass. Signs are a problem
for the voltage across the coil. The correct sign for any sign
convention is the sign which leads to energy conservation. In
the linear graph method, the voltage across the coil has the
same sign as the change in the current through the coil.


v12 = L

di
dt

(5.45)

The two differential equations above reduce to a relationship


between the time rate of change of the flux linkage and the
resulting voltage, v12 .

5.7 Electrical System Elements

293

d
= v12
dt

(5.46)

Equating the definition of flux linkage and the relationship


between flux linkage and inductance


= N = Li

(5.47)

yields an expression for the magnetic flux in terms of the


inductance of the coil, the current through the coil, and the
number of turns of the coil


Li
N

magnetomotive force mmf = Ni

(5.49)

The relationship between the strength of an externally applied magnetic field and the magnetic flux in the core of an
inductor (Eq.5.50) is analogous to the relationship between
the voltage applied to an electrical resistance and the current
through the resistance


Ni mmf = R

(5.50)

where R is reluctance which is a measure of the difficulty


of creating a magnetic field.
Magnetic reluctance varies with geometry, in the same
manner as electrical resistance varies with geometry,
Eq.5.51. Magnetic permeability is analogous to electrical
conductivity.


R =

Length
Area

(5.51)

Using mmf Ni = R yields


R =

b
L

N2
L

And using the definition of magnetic reluctance, Eq.5.51,


yields
L Length
= 2
N Area
This expression is useful in energy methods to calculate the
force, or torque, created by an electric motor.

iL

iL

Fig. 5.43a Schematic symbol for an air core inductor. bSchematic


symbol for an iron core inductor
Table 5.3 Mathematic analogy
between fluid inertance and
electrical inductance

(5.48)

The product of the current i passing through a coil and the


number of turns in the coil is the strength of the externally
applied magnetic field, and is known as the magnetomotive
force, abbreviated mmf.


Fluid
inertance
p12 = I

EI =

dQI
dt

1 2
IQI
2

Electrical
inductance
v12 = L

EL =

diL
dt

1 2
LiL
2

The analogy between electrical inductance and fluid inertance is very helpful, Table5.3.
The fluid inertance element represents the kinetic energy
storage, which results from the motion of fluid. A net pressure force is needed to accelerate or decelerate a mass of
fluid. Time is needed to change the amount of stored kinetic
energy, thereby changing the energy storage variable velocity. Electrical inductance is the magnetic energy storage,
which results from the motion of charge carriers (current).
A voltage is needed to increase or decrease a current. Time
is needed to change the amount of magnetic energy stored,
which changes the energy storage variable current. A typical
inductor is a coil of high conductivity copper wire wrapped
around a ferromagnetic core, such as iron, electrical steel,
or ferrite, which acts to enhance the magnetic field.
Inductance is counter-intuitive, unless an inductor is
viewed as an energy storage element. In steady-state, a coil
of copper wire has little resistance to the flow of current. In
the case of an ideal inductance, the coil has zero resistance
to the flow of current in steady-state. There is transient or
dynamic resistance which represents the flow of energy into
or out of the magnetic field created by the current. A voltage applied to a coil cannot instantaneously create a current
because the current through the coil is the energy storage element. A voltage applied to a coil leads to a rate of change in
the current through the coil, just as a force applied to a mass
leads to a rate of change in the velocity of the mass.

5.7.4 Electrical Systems


Schematics of electric circuits are models of the circuit. All
components of every circuit possess resistance, capacitance,
and inductance. The energetic attribute which dominates a
circuit component is the property we assign the component.

5 Fluid, Electrical, and Thermal Systems

294

V1

C
g

V2

Capacitance, C

2
+

v(t)

C
g

Resistance R1
Voltage source
Coil around ferrite core with
Induction L and Resistance R 2

Fig. 5.44a Electrical engineering convention, bEquivalent circuit


with voltage source
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
circuit are the resistance and inductance of the conductors
connecting the components. Occasionally, a designer errs
and uses too small a wire, or trace, on a printed circuit
board (PCB). The omission inadvertently creates a resistance
which affects the performance of the circuit.
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
show a portion of a circuit without the source, which renders it an incomplete circuit, Fig.5.44. The input variable is
identified as the voltage difference between the input node
and ground. The corresponding circuit includes the voltage
source.
Linear graphs are drawn, when working with an electrical
system for which there is no schematic. The linear graphs
for a coil that possesses significant inductance and resistance
must include a fictitious node, with no physical location, to
divide the voltage across the terminals of the coils between
the resistance and inductance properties, Figs.5.45 and 5.46.

5.8 Thermal Systems


Thermal systems do not fit the set of analogies used for
mechanical, electrical, and fluid systems. Thermal systems
have only one type of energy storage element, thermal capacitance. Also, power in a thermal system is not the product
of two variables. Heat flow rate, q, is both power and the
through variable in a thermal system.
Heat is energy. Heat exists in two very different forms. In
matter, heat is vibrational kinetic energy of atoms and molecules about their mean position. Heat is also transmitted
through matter or a vacuum as photons of electromagnetic
radiation in the infrared wave lengths.
As students of thermodynamics know, a flow of entropy
accompanies a flow of heat. Thermal systems are of great
importance in mechanical engineering due to our continued
reliance on heat engines as our prime movers. We will limit

our perspective to what is often referred to as thermal management, which is the removal of waste heat from machines
and devices. Thermal management is an essential aspect of
machine design, because a temperature increase due to the
accumulation of energy dissipated as heat limits the operating range of a device.

5.8.1 Thermal Power Variables


In a static system, the difference in temperature between two
locations determines the flow of heat. The across variable of
a thermal system is temperature, given the symbol, . Temperature is analogous to other across variables. If every location in a system is at the same temperature, then there is no
heat flow in the system. As noted above, the through variable
is heat flow rate, q.

5.8.1.1Analogies and Their Limits for Thermal,


Fluid, and Electrical Systems
The analogies across thermal, fluid, and electrical systems
are limited but useful. Heat flows in the direction of decreasing temperature. Electric current flows through a resistor in the direction of decreasing voltage, and is a potential driven flow, where the potential is electrical potential
or voltage. The strongest analogy to a thermal system is a
fluid system with flow velocity low enough, that its kinetic
energy hence inertance is negligible. A low velocity flow
of fluid through soil or a filter is a seepage flow. Seepage
flows are pressure driven, from high pressure to low pressure. Temperature and pressure both have an absolute zero
reference. Each uses a common, or human scale, reference:
zero Celsius and atmospheric pressure.
The analogies between the fluid and charge flow and the
flow of heat break down, when fluid velocity or amount of
current is large enough to store energy in either inertance or
inductance, which has no analogy in a thermal system. There
is no inertia in heat flow. Heat flow can start instantaneously, because the motion involved is on the atomic level.

5.8 Thermal Systems

295

Fig. 5.46a Electrical system


showing voltage nodes. A fictitious node is added to the coil to
divide the voltage drops due to
resistance and inductance, which
occur over the length of the coil.
b Linear graph of the electrical
system

Capacitance, C

R2

R1

v(t)

Resistance R1
Voltage source

2
Coil around ferrite core with
Induction L and Resistance R 2

We shall see that thermal capacitance and fluid capacitance have similar restrictions due to their direct reference to
their ground. Charge, fluid, and heat are stored in capacitances. However, the energy stored in an electrical or fluid
system is calculated as the square of the voltage or pressure,
respectively. The quantity of heat stored in a thermal capacitance is proportional to the temperature, not to the temperature squared. As such, thermal capacitance is mathematically analogous to storing energy, by elevating mass against
gravity.

5.8.2Modes of Heat Transfer and Their


Corresponding Thermal Resistances
Heat is energy. Consequently, heat flows. It does so spontaneously, in the direction of decreasing temperature. The
movement of heat is known as heat transfer. There are
three modes of heat transfer: conduction, convection, and
radiation.

5.8.2.1Conduction
Conductive heat transfer is the movement of heat by the
transfer of atomic or molecular kinetic energy between
neighbors in a static system. Fouriers law of heat conduction, Eq.5.52, describes heat flow:


q = k

d
dx

(5.52)

where is temperature, and k is thermal conductivity of the


material.
Conductive heat transfer usually refers to heat transfer
through a solid body, such as a cast iron frying pan or an
aluminum bar. The material need not be solid for conductive
heat transfer to occur, but it must remain static, i.e., motionless, at least on a small scale. Conductive heat transfer is
the movement of heat without the movement of matter. Heat
conduction is present in a liquid in a Lagrangian reference
frame, moving with the fluid. Conductive heat transfer completes the distribution of heat, analogous to diffusion completing the distribution of matter during the mixing of fluids.

The thermal conductivity of a single, homogeneous material is a specific property, which describes the rate of heat
flow q through a unit area normal to the heat flow path, per
unit length along the path for a unit temperature gradient,
-d/dx, in the direction of the heat flow. The conversion from
US customary units to SI is:
Thermal conductivity k : 1

Btu ft
W
= 1.730
hr ft 2 o F
m oK

(Conversion between US and SI units is complicated by a


slight difference in the definition of a Btu. The difference is
due to the variation in specific heat of water with temperature. The above conversion uses the US definition of a Btu.)
The thermal conductivity of all materials is a function
of temperature. Hence, thermal conductivity is a non-linear
property. It is linearized by modeling the thermal conductivity of a material as constant, using the thermal conductivity
for the expected average temperature.
A heat transfer coefficientis the rate of heat flow q through
a unit area per unit time.
Coefficient of heat transfer cp : 1

Btu
W
= 5.674 2
2 o
hr ft F
m K

where the subscript, p, stands for constant pressure.


In the case of conductive heat transfer, the heat transfer
coefficient, hc, is calculated, by multiplying the thermal
conductivity of the material by the length of the heat flow
path through the material. When heat passes through multiple materials in contact with one another, the overall heat
transfer coefficient is calculated, by modeling the heat path
as thermal conductivities in series, or, more conveniently,
as thermal resistances in series, as we shall in Sect.5.8.31,
Fig.5.47.

5.8.2.2Convection
Convective heat transfer is the movement of heat with the
movement of a fluid. Convective heat transfer combines
conduction of heat through fluid, and the movement of heated fluid. The division between the two mechanisms depends
on the thermal conductivity of the fluid, the heat capacity of

5 Fluid, Electrical, and Thermal Systems

296
Perfect Insulation
No heat flow

Heat Flow
Rate

Material 1

Material 2

Fig. 5.47 Heat conduction through two materials with different thermal conductivities. The thermal conductivity of Material 1 is less than
that of Material 2

the fluid, mixing of the flow, and the speed of the flow. If the
flow is turbulent, as is assumed for the flow of fluids in dynamic systems, then the mixing of heated fluid is thorough,
and the temperature is uniform over a cross-section in the
pipe, tube or hose, at a distance from where the heat entered
the fluid.
Convective heat transfer is classified as either free or
forced convection. Free convention occurs when the motion of the fluid against the higher temperature surface is due
to temperature dependence of the fluids density. Generally,
fluids become less dense with increasing temperature, as the
increased kinetic energy of their atoms or molecules increases the mean distance between them. The buoyancy of regions
of higher-temperature fluid causes it to float. The convective
motion allows the liquid at a heated surface to be replaced by
denser, colder fluid. Convective cells can be seen in water
being heated in a pan. Free convection is also noticeable in
the early morning, as the sun heats moist surfaces. The rising
mist is free convection.
Forced convection occurs when a pump or fan forces or
drives fluid motion against a heated surface. Early, coal-fired
central heating systems often relied on free convective flow.
Modern central heating systems have forced convection of
either heated air or water. Forced convection is clearly more
involved than free convection. The expense of the additional
hardware is offset by the increased efficiency of heat transfer
Fig. 5.48 Heat flow from fluid
1 through solid materials 1 and
2 into fluid 2. The heat transfer
through the thermal boundary layers in fluids 1 and 2 is
described by the film coefficients
for the two fluidsolid interfaces

from a solid surface to the fluid, due to the lower temperature


of the fluid against the solid surface. This is a familiar phenomenon to those who heat water on a stove. A pan or kettle
which has been gently tipped back and forth to agitate the
water heats faster than one left undisturbed. Increasing the
temperature difference between the heated surface and the
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
or pump, and the higher speed of the flow. Turbulent flow
enhances convective heat transfer, by mixing the fluid normal to the heated surface. The pressure gradient moves the
flow parallel to the heated surface, thereby reducing the temperature of the fluid on the heated surface.
Because the heat transfer from the solid to the liquid, by
definition, occurs in fluid that is in contact with the solid
surface, the heat transfer coefficient, hf , is called a film
coefficient, and applies only in a thermal boundary layer at
the solid-fluid interface, Fig.5.48.
Convective heat transfer in machine design usually occurs in a series of subsystems. For example, the power dissipated as heat in an automatic automobile transmission is
transferred from the transmission to the automobiles radiator by tubes, which connect to a coil within the tank of the radiator. The heat is then transferred to the water-based engine
coolant in the radiator. The heated water is pumped through
the tubes of the radiator, where it is transferred to air, that is
passed over the fins of the radiator by an electrically driven
fan.
Heat transfer in automotive design has always been important, and has become more so, as the engine compartments become smaller and more crowded, and the temperature within increases.

5.8.2.3Radiation
The heat radiated from a blackbody, an ideal radiator, with
surface area A through a hemisphere, qhemisphere, is described
by the StefanBoltzmann Law:


qhemisphere = A 4
Perfect Insulation
No heat flow

1
q

(5.53)

Solid
Material 2

Heat Flow
Rate
Fluid 1
Thermal
Boundary
Layer

Solid
Material 1

4
Fluid 2

Thermal
Boundary
Layer

5.8 Thermal Systems

297

Fig. 5.49 Thermal system consisting of a layer of stainless steel


and a layer of aluminum

Perfect Insulation
No heat flow

1
q
Heat Flow
Rate

Stainless
Steel

Surface Area
A = 48 ft 2

where T is the absolute temperature of the surface in degrees kelvin, and is the StefanBoltzmann constant,
= 5.67 108 W/(m 2 K 4 ).
The StefanBoltzmann law describes the radiation of heat
from a surface. That surface will also receive radiated heat
from other surfaces. The net radiative heat transfer between
two surfaces at different temperatures is the difference between the heated radiated from the surface, and the heat radiated to the surface, which was absorbed by the surface and
not reflected. The amount radiated will, in general, differ
from the ideal blackbody, and is corrected by the emissivity, , of a surface. Likewise, the fraction of the radiation
striking the surface which is absorbed will differ, and is corrected by the absorbance, .

5.8.2.4 Phase Change


Phase changes either store or liberate latent heat. Boiling
and evaporative cooling are the most common mechanisms
of phase change. Phase change of the working fluid in a hydraulic system, such as the boiling of brake fuel, leads to catastrophic failure. The brake components must be designed to
limit temperature rise in brake fluid.

5.8.3 Thermal System Elements


Electric circuit calculations are performed with electrical resistance, rather than its inverse, electrical conductance. Similarly, thermal system calculations are performed with thermal resistance, rather than its inverse, thermal conductivity. The single energy storage mode is thermal capacitance,
which must be referenced to ground temperature.

5.8.3.1 Thermal Resistance


The elemental equation for thermal resistance has the same
form as electrical and fluid resistance, potential=resistance
times flow rate:
v12 = RiR

and

p12 = R fluid QR fluid

12 = Rthermal qRthermal

(5.54)

Aluminum

2 in

0.5 in

where is temperature, q is heat flow rate, both the through


variable and power, and Rthermal is the thermal resistance.
Thermal resistance is calculated as the inverse of the
product of the heat transfer coefficient and the cross-sectional area of the heat flow path.


Rthermal =

L
k A

(5.55)

where k is thermal conductivity, L is the length of the heat


flow path, and A is the cross-sectional area of the heat flow
path.
Thermal resistances combine in the same way that electrical and fluid resistances combine, Sect.5.9.1. Resistances in
series sum to yield the equivalent thermal resistance. Resistances in parallel sum as inverses to yield the inverse of the
equivalent resistance.
Example: Figure5.49 shows a thermal system, consisting of a half-inch thick layer of stainless steel and a 2-in.
thick layer of aluminum. The area of the heat flow is 48ft2.
Stainless steel and aluminum have the following thermal
conductivity, k, at 212F:
k SS = 10

Btu
Btu
and k Al = 119
o
hr ft F
hr ft o F

Calculate the thermal resistance of each layer and the single


equivalent thermal resistance.
We will first convert to SI as standard practice. Electrical
units are SI, and problems involving systems with kinetic
energy storage are easier to work in SI. If the results must
be expressed in US Customary Units, we will convert back
at the end.
W
Btu
W
m oK
= 10
= 17.3
Btuft
hr ft o F
m oK
1
hr ft 2 o F
1.730

k SS

5 Fluid, Electrical, and Thermal Systems

298

W
Btu
W
m oK
k Al = 119
= 206
o
Btu ft
hr ft F
m oK
1
hr ft 2 o F
1.730

These thermal conductivities are specific conductivities for


an area of the heat path equal to the unit length squared, i.e.,
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
in either oF or oK.
The heat path lengths in meters are:
1 cm 1 m
LSS = 0.5 in
= 0.0127 m
2.54 in 100 cm
1 cm 1 m
LAl = 2.0 in
= 0.0508 m
2.54 in 100 cm

144 in 2 2.54 cm 1 m
A = 48 ft 2
= 4.459 m 2
1 ft 2 1 in 100 cm
The thermal resistances are:
RSS =

RAl =

o
0.0127 m
K
= 1.65 10 4
W
W
17.3
4.459 m 2
m oK
o
0.0508 m
K
= 5.53 10 5
W
W
2
206
4.459 m
m oK

The single equivalent thermal resistance for the thermal resistances, RSS and RAl, in series is the sum of those resistances.
Requiv = RSS + RAl = 1.65 10 4

o
K
K
+ 5.53 10 5
W
W

Requiv = 2.20 10 4

K
W

5.8.3.2 Thermal Capacitance


The specific heat of a material is defined as the amount of
heat required to raise a unit mass of the material by one degree:


cp = 1

Btu
kJ
= 4.184
o
lb m F
kg o K

Cp = M cp

(5.56)

where the subscript, p, stands for constant pressure. The specific heats of solids and fluids are functions of their tempera-

(5.57)

The heat capacity of a material, defined as the amount of


heat required to raise the temperature of a mass of the material one degree, is measured at either constant pressure or
constant volume. Therefore, the amount of heat stored in the
mass is the product of the change in temperature and the heat
capacity of the mass, Eq. 5.58.


The cross-sectional area of the heat path is:


2

ture, making the property non-linear. We customarily linearize the property, by using a representative value from within
the operating range. The heat capacity of an object is the
product of its specific heat, cp, times its mass, M:

(5.58)

EC = C p
p

Note the energy stored in a thermal capacitance is proportional to the change in temperature, not the change in the
temperature squared. The reason is that heat flow is power.
In all other types of systems, power is the product of the two
power variables of that system. In the energy storage equations, one power variable has been eliminated by an expression written in terms of the other, resulting in that variable
being squared.
The heat capacity of a heterogeneous object is the sum of
the heat capacities of the masses of materials, which comprise the object.

5.8.3.3Heat Sources: Friction, Fluid Shear,


and Electrical Resistance
Mechanical energy is dissipated as heat by friction, either
viscous friction due to the shear of fluids, or dry Coulomb
friction between two solid surfaces in contact. The power,
which flows into a translational or rotational damper, is dissipated as heat by viscous friction. It is important to clarify
the phrase, flows into, and distinguish it from the similar
phrase, flows through. Flows into means flows into
and is lost as heat. The power, which flows into a damper
and is lost as heat, is the product of the two power variables
of the damper. It is the product of the force, or torque, acting through the damper and the velocity difference across
the damper. The power supplied by the force source and
the power dissipated in dampers, b1 and b2, in the system
shown in Fig.5.50 are Psource = F (t ) v1g , Pb = Fb v12, and
Pb = Fb v2 g . The power, which flows through damper b1,
is the product of the force acting through the damper, and
the lesser of the velocities of the nodes at either end of the
damper, v2g, in this case.
1

Psource Pb = Pb
1

F (t ) v1g Fb1 v12 = Fb2 v2 g

Conventional automobile brakes dissipate power as heat


by dry friction. Dry friction is more difficult to include in

5.8 Thermal Systems

299

b1

teslas

b2

F(t)
g

The energy dissipated


as heat each cycle equals
the area of the hysteresis loop
amp.turns

H ______
meter

Fig. 5.51 Magnetic hysteresis dissipates magnetic energy as heat

Fig. 5.50 Mechanical system which acts as a heat source

dynamic models, because dry friction is a non-linear phenomenon. The friction force must reverse direction, when
the relative motion reverses direction. Modeling dry friction
requires numerical methods. An algorithm for dry friction is
presented in Chap.8.
Fluid energy is dissipated as heat, due to viscous shear.
Although all real fluids have viscosity, the effect of viscosity
on the flow of fluids in a dynamic model is represented by
fluid resistances. All power which flows into a fluid resistance is dissipated as heat.
Electrical energy is dissipated as heat, by electrical resistance. All power which flows into an electrical resistor is dissipated as heat. Magnetic hysteresis is the energy lost during
the cycle of creating and destroying a magnetic field in a
ferromagnetic material, such as the iron of an electric motor.
It is a significant source of heating in many motors. Magnetic
hysteresis is commonly included with the resistive heating
of motors, because both are a function of the current flow
through the motor.

5.8.3.4Heat Sources: Magnetic and Mechanical


Hysteresis
Hysteresis occurs, when the paths followed during loading and unloading of a system are not the same. Hysteresis
occurs in mechanical systems, when there is plastic deformation of a component during loading. The energy used to
create plastic deformation is dissipated as heat, and is not
recovered, when the component is unloaded. Hysteresis
also occurs when a magnetic field is created in a ferromagnetic material, such as iron, nickel, cobalt, and their alloys,
Fig. 5.51. Some of the work done rearranging the microscopic magnetic domains of the material, as the magnetic
moment of the atoms are aligned with the externally applied
magnetic field, is not recovered and is lost as heat. Magnetic
hysteresis is of great practical importance, because electrical machinery, i.e., electric motors, experience at least one
cycle of magnetic hysteresis per revolution. The power of
an electric motor is limited by its maximum current. The
maximum current, in turn, is limited by the temperature rise
of the motor in operation. Reducing the amount of energy

lost in magnetic hysteresis allows more current to be passed


through the motors windings.
A second phenomenon which heats an electric motor is
an eddy current. Eddy currents are induced in the iron and
steel of the motors rotor as it moves through the magnetic
field of the stator. The resistance in the rotor causes the eddy
currents to dissipate, heating the rotor. The rotors of many
electric motors are constructed of stacks of sheet steel stampings which are coated or painted to insulate the sheets from
one another, in order to reduce the length of a possible eddy
current path.

5.8.4 Thermal Systems


We will consider a system with thermal resistance and thermal capacitance, for the purpose of modeling heat diffusion.
Purely resistive thermal system models are used to estimate
the heat flow rates and temperatures in steady-state. In order
to model the transient period during which the system heats
up, thermal capacitances must be included to store heat.
Heat diffusion is the term which describes the transient
portion of heat flow. A materials property of thermal diffusivity increases with its thermal conductivity, and decreases with its specific heat. Increased thermal conductivity
increases the heat flow rate through the material. Increased
specific heat requires more heat to raise the materials temperature.
We will model the stainless steel and aluminum thermal
system of Sect.5.8.3.1 to create a linear graph to calculate
transient response of temperature. In order to model the
thermal capacitance of the metals, we will divide the aluminum into regions, one-half inch in thickness. The thermal
capacitance will be calculated for the entire mass of the region, using the temperature of the in-center of the region,
Fig.5.52. Heat will flow through thermal resistances equal
to the thermal resistance of the thickness of the regions, a
quarter of an inch.
The thermal resistance for half of the thickness of the stainless steel layer is half of what we calculated in Sect.5.8.3.1.

5 Fluid, Electrical, and Thermal Systems

300
Fig. 5.52 Thermal system divided into five regions, each with a
thermal capacitance between the
center temperature node of the
region and ground temperature

Perfect Insulation
No heat flow

q in

Heat Flow
Rate

1
0.0127 m
5 K
=
= 8.23 10
W
2
W
2
17.3 m o K 4.459 m

lb
Btu
and SS = 488 m3
lb m o F
ft

The specific heat and density of aluminum are:


lb
Btu
= 0.214 o
and Al = 169 m3
lb m F
ft

Convert these values from US customary units to SI:


kJ

4.184
Btu
kJ
kg o K
= 0.11 o
= 0.46
Btu
lb m F
kg o K

lb o F
m
3

SS = 488

c pAl

lb m 0.454 kg 1 ft 1 in
kg
= 7,820 3
ft 3 lb m 12 in 0.0254 m
m

kJ

4.184

Btu
kJ
kg o K
= 0.214 o
= 0.895
Btu
lb m F
kg
oK

1 lb o F
m

RAl

RAl
8

2 in

The specific heat and density of stainless steel (18 Cr, 8 Ni)
are:

c pSS

0.5 in

1
0.0508 m
6 K
RAl =
= 6.91 10
W
8
W
2
206 m o K 4.459 m

c pAl

Aluminum

Similarly, the resistance of one of the eight 2-in. thick aluminum layers is

Stainless
Steel

Thermal resistance scales with the length of the heat flow


path, all else being equal.

c pSS = 0.11

RAl RAl

RAl

RAl

RAl
3

Surface Area
A = 48 ft 2

RSS1 = RSS2

RSS

RSS

Al = 169

RAl
10

11
qout

lb m 0.454 kg 1 ft 1in
kg
= 2, 710 3
ft 3 lb m 12 in 0.0254 m
m

The thermal capacitance, Cp, of the half-inch thick layer of


stainless steel equals the product of its mass, M, and specific
heat, cp.
0.0254 m
kg
4.459 m 2 7,820 3 = 443 kg
M SS = 0.5 in

m
1 in
C pSS = Mc pSS = 443 kg 0.46

kJ
kg o K

C pSS = 204

kJ
K

The thermal capacitance of a half-inch thick region of the


aluminum is
0.0254 m
kg
4.459 m 2 2, 710 3 = 153 kg
M Al = 0.5 in

m
1 in
kJ
kJ
 C pAl = Mc pAl = 153 kg 0.895 kg o K C pAl = 137 o K
The linear graph of this model is shown in Fig.5.53.
The linear graph is intimidating, but it can be easily reduced by the state-space method introduced in Chap.7. The
state-space method can deal with multiple inputs. There are
two temperature sources in the linear graph to specify the
surface temperatures on ends. The reduction of this linear
graph to a system of state equations and their solution comprise Problem 7.20 of Chap.7.

5.9Equivalent Elements in Fluid, Electrical,


and Thermal Systems
When two or more like elements are in series or in parallel,
we can simplify the systems model, by combining the like
elements into a single equivalent element. The equivalent element will store or dissipate power at the same rate, as the
sum of the power stored or dissipated by the original ele-

5.9 Equivalent Elements in Fluid, Electrical, and Thermal Systems


Fig. 5.53 Linear graph of the
stainless steel and aluminum
thermal system. The linear graph
is split between nodes 6 and 7 for
this figure

R ss1

301

Rss 2

R al 1

Css

1(t)

R al2

R al 3

Ral 4

C al1

Cal 2

Ral 4

R al 5

Ral 6

R al 7

10

Cal 3

Cal 2

Ral 8

11

Cal 4

2(t)
g

Fig. 5.54a Resistances in series.


b Equivalent resistance

R1

1
Fig. 5.55a Resistances in
parallel. b Equivalent resistance

R2

R equiv
3

R1
1

R equiv
1

R2
ments. When elements in series or in parallel are combined
and replaced by one equivalent element, the sum of either
the across variables or the through variables in the original
elements will equal the corresponding power variable in
the equivalent element. Elements in series have the same
through variable. Conversely, elements in parallel have the
same across variable.
The values of the equivalent parameters are determined
by writing the continuity, compatibility, and elemental equations for the portion of the linear graph, which contains the
similar elements to be replaced. The set of equations is reduced to the form of the elemental equation, written in terms
of the through and across variables for the equivalent element.

5.9.1 Fluid, Electrical, or Thermal Resistances


Resistances in series sum to yield the equivalent single resistance. Resistances in parallel sum as inverses to yield the
inverse of the single equivalent resistance.

5.9.1.1 Fluid, Electrical, or Resistances In Series


Elements in series have the same through variable, Fig.5.54.
The sum of their across variables equals the across variable

of the equivalent element. We will derive the equivalent


electrical resistance. The equivalent resistance for the corresponding fluid resistances is calculated in the same way.
Continuity iR1 = iR2 = iRequiv
Compatibility v12 + v23 = v13
Elements v12 = R1iR1

v23 = R2 iR2

v13 = Requiv iRequiv

Reduction
v12 + v23 = v13

R1iRequiv + R2 iRequiv = Requiv iRequiv

R1iR1 + R2 iR2 = Requiv iRequiv

( R1 + R2 ) iR

equiv

= Requiv iRequiv

Requiv = R1 + R2

5.9.1.2Fluid, Electrical, or Thermal Resistances In


Parallel
Elements in parallel have the same across variable, Fig.5.55.
The sum of parallel elements through variables equals the
through variable of the equivalent element.

5 Fluid, Electrical, and Thermal Systems

302
Fig. 5.56a Fluid and thermal
capacitances must be referenced to a ground pressure or
temperature. Two fluid or two
thermal capacitances in parallel.
b Single equivalent fluid or
thermal capacitance. c Electrical capacitances do not need a
ground reference. Two electrical
capacitances in parallel. d Single
equivalent electrical capacitance

C1

Continuity iR1 + iR2 = iRequiv


Compatibility v12 = v12
Elements v12 = R1iR1

v12 = R2 iR2

v12 = Requiv iRequiv

Reduction
iR1 + iR2 = iRequiv
1
1
iRequiv = + v12
R1 R2
iRequiv =

R1 + R2
v12
R1 R2

Requiv =

v12 v12
+
= iRequiv
R1 R2
R
R
iRequiv = 2 + 1 v12
R1 R2 R1 R2

v12 =

Cequiv

C2

R1 R2
iR
R1 + R2 equiv

R1 R2
R1 + R2

5.9.2 Fluid and Electrical Capacitance


Fluid and electrical capacitance both store energy as a function of their across variables, fluid pressure, and electrical
voltage. However, fundamental physical differences between fluid pressure and electrical voltage, as well as those
between volume flow rate and electric current, limit the
analogy between them. The pressure in a fluid capacitance
must be measured relative to a minimum ground pressure,
either atmospheric pressure or a complete vacuum. Substantial negative pressures can be created on a small scale
under careful laboratory conditions, by creating tension in
liquids in capillary tubes. However, negative pressure below
an absolute vacuum does not exist in practice, except for
the melt strength of polymers which have some tensile
strength when fluid. Low-molecular-weight fluids cavitate,
or boil, before they reach negative pressure.
Volume flow rate and current (electric charge flow rate)
are also fundamentally different from each other. While there
are both positive and negative charges, there is only positive volume flow rate. Current appears to flow through an

C1

Cequiv

C2

electrical capacitor. In fact, charge carriers (electrons and


positive ions) do not pass through the dielectric (insulator)
separating the oppositely charged plates of an electrical capacitor. However, the opposite polarity charge carriers flow
into the capacitor on opposite sides, moving in opposite directions. Fluid capacitances have a single connection which
serves as both the inlet and outlet ports, depending on the
direction of the flow. Fluid cannot flow through a fluid capacitance without violating the physical model, by overtopping a standpipe or rupturing a fluid accumulator.
The net result of the physical differences between fluid
and electrical capacitances is that energetic elements in circuits and linear graphs have different constraints. Electrical capacitance is more flexible, and can be placed anywhere in a circuit, except in direct connection with a voltage
source. Electrical capacitances do not need to be referenced
to ground. Consequently, electrical capacitances can be connected in series. Fluid capacitances must be referenced to
ground. Hence, fluid capacitances cannot be connected in
series.
Thermal capacitance is similar to fluid capacitance, in that
it is referenced to ground temperature. Consequently, neither
thermal capacitances nor fluid capacitances can be connected
in series. Attempting to connect a thermal capacitance in series simply results in a larger heterogeneous thermal capacitance. Another difference between thermal and fluid capacitance is the energy equation of each. Energy stored in a fluid
capacitance is proportional to the square of the pressure in
the capacitor. Heat stored in a thermal capacitance is linearly
proportional to an ideal thermal capacitance.

5.9.2.1Fluid, Electrical, and Thermal Capacitances


In Parallel
Elements in parallel have the same across variable, Fig.5.56.
The sum of their through variables equals the through variable of the equivalent element.
Continuity iC1 + iC2 = iCequiv
Compatibility v12 = v12

5.9 Equivalent Elements in Fluid, Electrical, and Thermal Systems


Fig. 5.57a Two electrical
capacitances in series. b Single
equivalent electrical capacitance

303

C1

C2

Fig. 5.58a Parallel fluid inertances, b Equivalent single fluid


inertance, c Parallel electrical
inductances, d Equivalent single
electrical inductance

C equiv
3

I1
1

I equiv

I2

L1
1

L equiv

L2

Elements iC1 = C1

dv12
dt

iC2 = C2

dv12
dt

iCequiv = Cequiv

dv12
dt

C + C2
1
= 1
Cequiv
C1C2

Cequiv =

C1C2
C1 + C2

Reduction
iC1 + iC2 = iCequiv

(C1 + C2 )

C1

dv12
dv
dv
+ C2 12 = Cequiv 12
dt
dt
dt

dv12
dv
= Cequiv 12
dt
dt

Cequiv = C1 + C2

5.9.2.2 Electrical Capacitances In Series


Elements in series have the same through variable. The sum
of their across variables equals the across variable of the
equivalent element, Fig.5.57.
Continuity iC1 = iC2 = iCequiv

dv12
dt

iC2 = C2

dv23
dt

iCequiv = Cequiv

dv13
dt

Reduction
v12 + v23 = v13
iC1
C1

iC2
C2

iCequiv

1
1
1
+
=
C1 C2 Cequiv

dv12 dv23 dv13


+
=
dt
dt
dt
iCequiv

Cequiv

1 C2 1 C1
1
+
=
C1 C2 C2 C1 Cequiv

5.9.3.1Fluid Inertance or Electrical Inductance


In Parallel
Elements in parallel have the same across variable. The sum
of their through variables equals the through variable of the
equivalent element, Fig.5.58.
Continuity iL1 + iL2 = iLequiv
Compatibility v12 = v12
Elements v12 = L1

Compatibility v12 + v23 = v13


Elements iC1 = C1

5.9.3 Fluid Inertance or Electrical Inductance

C1

iCequiv
C2

iCequiv
Cequiv

1 C2 1 C1
1
+
=
C1 C2 C2 C1 Cequiv

C2
C
1
+ 1 =
C1C2 C1C2 Cequiv

diL1

v12 = L2

dt

diL2

v12 = Lequiv

dt

diLequiv
dt

Reduction
iL1 + iL2 = iLequiv
v12 v12
v
+
= 12
L1 L2 Lequiv
1
Lequiv

diL1

dt

L2
L
+ 1
L1 L2 L1 L2
Lequiv =

diL2
dt

diLequiv
dt

1 1
1
+
=
L1 L2 Lequiv
1
Lequiv

L1 L2
L1 + L2

L1 + L2
L1 L2

5 Fluid, Electrical, and Thermal Systems

304
Fig. 5.59a Fluid inertances in
series. b Equivalent single fluid
inertance. c Electrical inductances in series. d Equivalent single
electrical inductance

I1

L1

Elements in series have the same through variable, Fig.5.59.


The sum of their across variables equals the across variable
of the equivalent element.
Continuity iL1 = iL2 = iLequiv
Compatibility v12 + v23 = v13
diL1
dt

v23 = L2

diL2

v13 = Lequiv

dt

diLequiv
dt

Reduction
v13 = v12 + v23
v13 = L1

diLequiv
dt

+ L2

diLequiv
dt

v13 = L1

diL1
dt

+ L2

diL2
dt

v13 = ( L1 + L2 )

diLequiv
dt

Lequiv = L1 + L2
Two or more fluid inertances in series combine similarly,
I equiv = I1 + I 2 .

Summary
The physical and mathematical analogies between electrical
and fluid systems are very strong, but not perfect. The physical and mathematical analogies across thermal, electrical,
and fluid systems are quite limited.
Fluid inertance is a helpful physical analogy for electrical inductance. The need to accelerate and decelerate a mass
of fluid, in order to store or remove kinetic energy, is a fact
that we have all experienced. Electrical inductance is a new
concept to many students. It is baffling that voltage is tempo-

I equiv

L2
2

5.9.4Fluid Inertances or Electrical Inductances


In Series

Elements v12 = L1

I2
2

L equiv
3

rarily established across a coil of low resistance wire, when


there is an attempt to change the current through the coil,
except when the current is recognized as storing energy in
a magnetic field. The storage or remove of energy requires
time.
Fluid, electrical, and thermal systems all have resistance
and capacitance. The mathematical analogy among the resistances in three types of physical systems is that the element
equation is expressed as a proportion between the power
variables of the system. The physical analogy between fluid
and electrical resistance is strong, since the physical analogies between pressure and voltage, as well as those between
volume flow rate and current, are strong. The physical analogy of dissipating heat due to friction or a friction-like phenomenon in a flow restriction describes both electrical and
fluid resistance. Thermal resistance is resistance to flow. Reducing the cross-section of the flow and lengthening the flow
path have the same effect as in electrical or fluid resistance.
However, thermal resistance cannot dissipate energy as heat,
because the energy in a thermal system is heat. Further, the
power cannot flow into a thermal resistance. Power can only
flow through a thermal resistance.
An electrical capacitance does not need to be referenced
to ground. The energy stored in an electrical capacitance is a
function of the voltage across the capacitor, regardless of the
voltages of the two plates. Other than a fluid systems capacitors reference to ground pressure, the physical analogies
between the electrical and fluid capacitors are strong. One
can imagine working to cram additional fluid into a fluid
accumulator, and working to cram additional charge onto
the plates of an electrical capacitor against every increase
in pressure or voltage. The mathematical analogy between
the fluid and electrical capacitor is perfect. Although one can
imagine trying to cram additional heat into a thermal capacitance, the mathematical analogy is absent, because energy
storage in a thermal capacitor is a linear function of temperature, not a function of temperature squared.

Problems

305

R1

Table 5.4 US Customary to SI Unit Conversions


1 in=2.54 cm=25.4 mm
1 MPa=145 psi
1 horsepower=746 W
1 gallon=3.79 L
1 Btu = 1, 055 J

v(t)
C

vent to
atmosphere

P(t)
Pump

R2

Fig. P5.2 RL circuit schematic

R1

R1

R2

Fluid
Reservoir

v(t)
Fig. P5.1 Fluid system schematic

R2

Problems

Fig. P5.3 RC circuit schematic

Reminders
1. Use nodal notation for the across variable drop, and indicate the positive direction of the through variable in the
direction of the drop in the across variable. Sources are
the exception. Sources raise the across variable in the
direction of the through variable flow.
2. Check units in terms of the power variables of the system,
not fundamental units or conventional units.
3. Although system dynamics calculations can be performed
using US customary units, the set of US units can lead to
significant errors, due to the mishandling of the gravitational constant. The units of mass moment of inertia are a
case in point. US customary units commonly used ounce
in2, lb in2, lb ft2, and slug ft2. It is easier to perform the
calculations in SI Units and then convert to US customary
units to present the results, if so required.
4. A dynamic system has only one characteristic equation,
independent of our choice of the output variable. Consequently, there is a time constant of the system, only one.
All power variables in the system vary at the same rate,
although their step responses will differ. Some variables
will grow, while others will decay, but they do so in synchronizativon.
5. Remember to convert from kPa and MPa to Pa for SI
units.
6. There are 1,000L in 1m3.
7. Engineers practicing in the United States must be able to
convert between US customary units and SI units without
resorting to a reference or the internet. Before you leave
school, commit to memory at least one conversion factor
for each unit you will work with. The most common conversion needed is length, i.e., 1 in = 2.54 cm = 25.4 mm ,
Table5.4.

Problem 5.1 A fluid system consisting of two fluid resistances, R1 = 100 MPa sec/m3 and R2 = 200 MPa sec/m3 , and a
fluid accumulator (capacitance), C = 6 10 8 m3 /Pa, is acted
upon by a pressure source, Fig. P5.1. Resistance R2 discharges to atmospheric pressure in the systems reservoir.
5.1.a Derive the system equations for:
i The volume flow rate from the pressure source.
ii The volume flow rate into the fluid accumulator.
iii The pressure in the fluid accumulator.
and check their units
The capacitor is de-energized at time, t = 0 . A step
change in pressure, P (t ) = 200 kPa us (t ) is applied at time,
t = 0.
5.1.b Solve the system equations of part a.
5.1.c Plot the responses using Mathcad or MATLAB
Problem 5.2An RL electric circuit where R1 = 10 ,
R2 = 20 , and L = 3 10 3 H is shown schematically in
Fig.P5.2.
5.2.a Derive the system equations for:
i The current through the inductor.
ii The voltage drop across the inductor.
iii The current through resistor R2.
and check their units.
The circuit is de-energized at time, t = 0 . At t = 0, a step
input voltage v(t ) = 48 VDC us (t ) to the circuit.
5.2.b Solve the system equations of part a.
5.2.c Plot the responses using Mathcad or MATLAB.
Problem 5.3 An electric circuit with R1 = 1 k , R2 = 2 k ,
C = 1 F and is shown schematically in Fig.P5.3.
5.3.a Derive the system equations for:

5 Fluid, Electrical, and Thermal Systems

306

vent to
atmosphere

R1

p(t)

Pump

R1

R2

I
v(t)

Fluid
Reservoir

Fig. P5.4a Fluid circuit schematic

p(t)
psi

R2

Fig. P5.5 RL circuit schematic

R1

2,000

1,000
0

v(t)
0

t, sec

R2

3
Fig. P5.6a RC circuit schematic

Fig. P5.4b Input pressure pulse

i The current through the capacitor.


ii The voltage across the capacitor.
iii The voltage across resistor R1.
and check their units.
The circuit is de-energized at time, t = 0 . A step input of
voltage v(t ) = 48 VDC us (t ) is applied to the circuit at time,
t = 0.
5.3.b Solve the system equations of part a.
5.3.c Plot the responses using Mathcad or MATLAB.
Problem 5.4The fluid system shown schematically in
Fig.5.4a consists of a fluid power unit modeled as a pressure source, two contractions modeled as fluid resistances, R1 = 100 MPa sec/m3 and R2 = 200 MPa sec/m3 ,
and a long run of piping modeled as a fluid inertance,
I = 40 106 kg/m 4 .
5.4.a Derive the system equations for:
i The volume flow rate from the pressure source.
ii The pressure difference across the fluid resistance,
R1.
iii The volume flow rate through the fluid inertance.
iv The pressure difference across the fluid inertance.
and check their units.
The system was de-energized before the pressure pulse
shown in Fig. 5.4b was applied to the system.
5.4.b Solve the system equations of part a.
5.4.c Plot the responses using Mathcad or MATLAB.
Problem 5.5An electric circuit with two resistors,
R1 = 10 and R2 = 20 , and an inductor, L = 3 10 3 H ,
is shown schematically in Fig. P5.5.
5.5.a Derive the system equations for:

i The current from the voltage source.


ii The voltage across resistor R1.
iii The current through the inductor.
iv The voltage across the inductor.
v The current through resistor R2.
and check their units.
The circuit is de-energized at time, t = 0 . At time, t=0,
a step input voltage v(t ) = 48 VDC us (t ) is applied to the
circuit.
5.5.b Solve the system equations of part a.
5.5.c Plot the responses using Mathcad or MATLAB.
Problem 5.6An electrical system consisting of a voltage
source, two resistors, R1 = 10 and R2 = 20 , and a capacitor, C = 1 10 6 F is shown in the schematic, Fig. P5.6a.
5.6.a Derive the system equations for:
i The current from the voltage source.
ii The voltage across resistor R1.
iii The current through the capacitor.
iv The voltage drop across the capacitor.
v The current through resistor R2.
and check their units.
The voltage applied to the system is plotted in Fig 5.6b.
5.6.b Solve the system equations of part a.
5.6.c Plot the responses using Mathcad or MATLAB.
Problem 5.7 An electric circuit with R1 = 1 k , R2 = 2 k ,
C = 1 F is shown schematically in Fig.P5.7.
5.7.a Derive the system equations for:
i The current through the capacitor.
ii The voltage across the capacitor.
iii The voltage across resistor R1.
and check their units.

Problems

307

R1

v(t), VDC
v2 = 36

R2
v1 = 12
0

v(t)

R3

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.
5.7.b Solve the system equations of part a.
5.7.c Plot the responses using Mathcad or MATLAB.
Problem 5.8 An electric circuit with R1 = 1 k , R2 = 2 k ,
C1 = 1 F and C2 = 2 F is shown schematically in Fig.P5.8.
5.8.a Derive the system equations for:
i The current through the capacitors.
ii The voltage across the capacitors.
iii The voltage across the resistors.
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.
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.
5.8.b Solve the system equations of part a.
5.8.c Plot the responses using Mathcad or MATLAB.
Problem 5.9 The electrical system shown in Fig. P5.9a consists of voltage source, a resistor with resistance, R1 = 100 ,
and a coil wound around a ferrite core. The coil has both
inductance, L = 3 10 3 H and resistance, R2 = 20 .
5.9.a Derive the system equations for:
i The current from the voltage source.
ii The voltage across resistance R1.
iii The current through the inductor.
iv The voltage across the inductor.
v The current through resistance R2.
and check their units.
The system was de-energized before the voltage pulse
shown in Fig.P5.9b was applied to the system.
5.9.b Solve the system equations of part a.
5.9.c Plot the responses using Mathcad or MATLAB.
Problem 5.10The fluid system shown schematically in
Fig.5.10a consists of a fluid power unit modeled as a pres-

R2

C1

v(t)

C2

Fig. P5.8 RC circuit schematic


Capacitance, C

Resistance R1

Voltage source
Coil around ferrite core with
Induction L and Resistance R 2

Fig. P5.9a Electrical system

v(t)
VDC

24
12
0

t, millisec

Fig. P5.9b Input voltage pulse

sure source, two contractions modeled as fluid resistances


R1 = 100 MPa sec/m3 and R2 = 200 MPa sec/m3, a long run
of piping modeled as a fluid inertance, I = 40 106 kg/m 4
and a fluid accumulator (capacitance), C = 6 10 8 m3 /Pa .

5 Fluid, Electrical, and Thermal Systems

308

C
P(t)

R1

Pump

vent to
atmosphere

R2

v(t)

Fluid
Reservoir

Fig. P5.11a RLC circuit schematic


Fig. P5.10a Fluid circuit schematic

v(t), VDC
p(t)
psi

v(0+) = 36

2,000

v(0-) = 12

1,000
0

t, sec

time

Fig. 5.11b Input voltage

Fig. P5.10b Input pressure pulse

5.10.a Derive the system equations for:


i The volume flow rate from the pressure source.
ii The pressure in the fluid accumulator.
iii The volume flow rate through the fluid inertance.
iv The pressure difference across the fluid inertance.
and check their units.
The system was de-energized before the pressure pulse
plotted in Fig.P5.10b was applied to the system.
5.10.b Solve the system equations of part a.
5.10.c Plot the responses using Mathcad or MATLAB.
Problem 5.11 An electrical system consisting of a voltage
source a resistor, R1 = 1 , a capacitor C = 1 10 6 F , and an
inductor, L = 3 10 3 H , is shown in the schematic P5.11a.
5.11.a Derive the system equations for:
i The current through the inductor.
ii The voltage across the inductor.
iii The current through the capacitor.
iv The voltage across the capacitor.
and check their units.
The system was in steady-state under the previously applied step input of 12 VDC when a step change was made at
time, t = 0, increasing the voltage to 36VDC, as shown in
Fig.5.11b.
5.11.b Solve the system equations of part a.
5.11.c Plot the responses using Mathcad or MATLAB
Problem 5.12 A firefighting system consists of a pump and
a 135ft. long, horizontal, circular cross-sectioned steel pipe,
terminated with a nozzle, as shown in Fig. P5.12. The pump

draws water from a reservoir at atmospheric pressure. The


pump is modeled as a pressure source in series with the internal resistance of the pump. The pressure of the source is
150psi. The maximum steady-state flow rate of the pump is
1250 gallons per minute when discharging directly to the atmosphere. The maximum steady-state flow rate of the pump
discharging through the pipe and nozzle is 800 gallons per
minute.
5.12.a What is the internal resistance of the pump?
5.12.b What is the internal diameter of the pipe in inches, if
the system reaches 70% of its steady-state flow in 2
sec after the nozzle valve is opened?
Problem 5.13An electric circuit consisting of a voltage
source, two resistors, R1 = 10 and R2 = 2 , a capacitor
C = 2 10 6 F, and an inductor, L = 0.015 H, is shown as the
schematic Fig.P5.13a.
5.13.a Derive the system equations for:
i The voltage across the capacitor.
ii The current through resistor R1.
iii The current through the inductor.
iv The voltage drop across the inductor.
v The voltage across resistor R2.
and check their units.
The system is initially de-energized before the voltage
pulse shown in Fig.P5.13b was applied.
5.13.b Solve the system equations.
5.13.c Plot the responses using Mathcad or MATLAB.
Problem 5.14 The electric circuit shown in Fig.P5.14 consists of a voltage source, a resistor, R = 1 , a capacitor,

Problems

309

Internal Resistance
of the Pump, R Internal

Fig. P5.12 Firefighting system

135 ft

p(t)

Pump

vent to
atmosphere

Nozzle
Resistance R Nozzle

Pipe, Inertance I

Fluid
Reservoir

L
v(t)

R1

R2

v(t)

Fig. P5.14 RLC circuit

Fig. P5.13a RLC circuit schematic

L1

10
5

v(t)
0
VDC
-5

L2
1

3
4
t, millisec

v(t)

C1
R

C2

-10
Fig. P5.13b Input voltage

C = 2 10 F, and coil with 1.4millihenrys of inductance


and 2 of resistance.
5.14.a Derive the system equations for:
i The voltage across the capacitor.
ii The current through resistor R.
iii The current through the inductor.
iv The voltage drop across the inductor.
and check their units.
The system is initially de-energized before a step input of
24 VDC was applied.
5.14.b Solve the system equations.
5.14.c Plot the responses using Mathcad or MATLAB.
Problem 5.15The electric circuit shown in the schematic, Fig.P5.15a, consists of a voltage source, a resistor,

Fig. P5.15a RLC circuit schematic

R = 10 , two capacitors, C1 = C2 = 2 10 6 F , and two inductors, L1 = L 2 = 0.015 H ,


5.15.a Derive the system equations for:
i The voltage across capacitor C2.
ii The current through resistor R.
iii The current through inductor L1.
iv The voltage drop across the inductors.
v The voltage across resistor R.
and check their units.
The system is initially de-energized before the voltage
pulse shown in Fig.P5.15b was applied.
5.15.b Solve the system equations.
5.15.c Plot the responses using Mathcad or MATLAB.
Problem 5.16A hydraulic system consists of a pump
modeled as a pressure source which draws fluid from a
reservoir vented to the atmosphere, two fluid resistances,

5 Fluid, Electrical, and Thermal Systems

310

100

C1

50

v(t)
0
VDC
-50

p(t)

3
t, millisec

R1

C2
R2

Pump

vent to
atmosphere

R3

Fluid
Reservoir

-100
Fig. P5.17a Fluid system schematic
Fig. P5.15b Input voltage

3,000

Fluid Power Unit Modeled as


Pressure Source p(t)
Breather Cap (vent to atmosphere)
High Pressure Output
Hydraulic Line
Resistance R1
Low Pressure Return

Fluid Accumulator
Capacitance C

Hydraulic Line
Inertance I and
Resistance R 2

Fig. P5.16 Fluid system

R1 = 100 MPa sec/m3 and R2 = 200 MPa sec/m3, and a


fluid accumulator (capacitor), C = 0.0040 m3 /MPa . The return line is 3/4in. I.D. and 40feet long. The density of the
hydraulic oil is approximately 875kg/m3.
5.16.a Derive the system equations for:
i The pressure in the fluid accumulator.
ii The volume flow rate into the fluid capacitor.
iii The volume flow rate through fluid resistance R2.
iv The volume flow rate from the pump.
and check their units.
5.16.b Calculate the fluid inertance of the return line
The system was running in steady-state with the pump
pressure of 1,800psi when, at time, t = 0, the pressure was
given a step increase to 3,000psi.
5.16.c Solve the system equations of part a and plot the
responses using Mathcad or MATLAB.
5.16.d The fluid reservoir of a fluid power unit is sized so
that the residence time, the average time that fluid
remains in the reservoir during operation, is two minutes, in order to degas the hydraulic fluid.
i Determine the capacity of the pump needed to
provide 125
% maximum flow calculated in
partc.

p(t)
psi

2,000

1,000

t, sec

Fig. P5.17b Input pressure pulse

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.
Problem 5.17The hydraulic system schematic shown
in Fig.5.17a consists of a pump modeled, as a pressure
source, which draws fluid from a reservoir vented to the
atmosphere, three fluid resistances, R1 = 100 MPa sec/m3 ,
R2 = 200 MPa sec/m3, and R3 = 300 MPa sec/m3, and two
fluid accumulators (capacitors), C1 = 0.04 m3 /MPa and
C2 = 0.05 m3 /MPa.
5.17.a Derive the system equations for:
i The pressure in the fluid accumulator, C1.
ii The volume flow rate into the fluid capacitor, C1.
iii The pressure in the fluid accumulator, C2.
iv The volume flow rate into the fluid capacitor, C2.
v The volume flow rate through fluid resistance R2.
vi The volume flow rate from the pump.
and check their units.
The system was de-energized, when it was given the pressure pulse plotted in Fig.P5.17b.
5.17.b Solve the system equations of part a and plot the
responses using Mathcad or MATLAB.
Problem 5.18 The fluid system shown in the Fig.P5.18a
consists of a pump, modeled as the pressure source, P(t),

Problems

311

Rcoil

C
R1

P(t)

vent to
atmosphere

Pump

v(t)

R2

L coil
1 F

Fluid
Reservoir

Fig. P5.19a RLC circuit schematic


Fig. P5.18a Fluid system schematic

Determine the unknown fluid resistance, R2 and capacitance, C. Report your results in SI units.

250
200

Q pump
gal
___
min

150
100
50
0

time, seconds

Fig. P5.18b Pump test data

two fluid resistances, R1 and R2, and fluid capacitance, C.


The pumps reservoir is vented to the atmosphere. The value
of the fluid resistance, R1 is known, R1 = 300 MPa sec/m3.
The values of the parameters R2 and C are unknown and to be
determined from a dynamic test. The pump was running in
steady-state with a pressure of 500psi when, at time t = 2.0
sec, the pressure was suddenly increased to 1,000psi.
The pumps pressure and volume flow rate are plotted,
Fig. P5.18b, and tabulated, TableP5.18, in US customary
units.

Table P5.18 Pump Test Data

Problem 5.19The RLC circuit shown schematically in


Fig.P5.19a consists of a capacitor in series with 2m of 22
AWG solid copper wire wound around a ferromagnetic core.
The resistance 22 AWG solid copper wire is 16.8ohms per
1,000ft. Note the coil has two energetic properties, resistance and induction. Induction enhanced by the presence of
a ferromagnetic material leads to energy loss, due hysteresis
of the magnetic field in the ferromagnetic material. Induction enhanced by ferromagnetic material is also non-linear.
The circuit was charged to 15 VDC. Then the connection
between the voltage source and the circuit and the end of the
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.
5.19.a Use the value of the resistance of the copper wire
above to determine the capacitance and inductance
of the circuit.
5.19.b Calculate the value of the apparent resistance, Rcoil,
from the observed response of the circuit. Determine
the capacitance and inductance of the circuit.
5.19.c How different are the values calculated in parts a and
b? If the difference is significant, what may account
for the discrepancy?

Time (sec)

Pump pressure (psi) QpumpGPM

Time (sec)

Pump pressure (psi)

QpumpGPM

0
1.0
2.0
2.1
2.2
2.3
2.4
2.5
2.6
2.7
2.8
2.9
3.0

500
500
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000

3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8
3.9
4.0
5.0
6.0
7.0

1000
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000
1000

123
119
115
112
110
107
105
103
102
101
94
92
91

45.5
45.5
228
211
196
183
171
161
152
145
138
132
127

5 Fluid, Electrical, and Thermal Systems

312
Fig. P5.19b The voltage across
the capacitor during the latter
portion of the discharge of the
RLC circuit

1.5

1.0

0.5

vcap ,VDC

-0.5

-1.0
0.0

Chapter 5 Appendix
Engineering Electromagnetics
A basic understanding of magnetic phenomena is essential
for mechanical engineers, given the importance of electric
motors in machine design. We will begin with the basics of
electromagnetics. With all physical phenomena, a realistic engineering model is a matter of degree. We will keep
our model of electromagnetics simple and ideal, so that it is
generally representative and applicable, but does not require
(much) vector calculus. In any case, the non-linearity and
uncertainty of the material properties, magnetic permeability
in this case, necessitates the use of a simple model, unless the
engineer has the need and the means to fully characterize the
materials he or she is working with.
We will introduce an engineering model created to exploit the similarity between a resistive electric circuit, and
the steady-state behavior of an electromagnet. We then investigate the non-linear magnetic properties of the ferromagnetic materials we use to enhance the strength of the magnetic fields we create for electric motors. We conclude this
chapter with the topic of flux linkage, which we will use in
energy method analyses of electric motors.

Electromagnetic Force
All magnetism is electromagnetism. The charge can be an
electron orbiting an atomic nucleus or an electric current.
The magnetic fields of permanent magnets are created by
the alignment of the individual magnetic fields of electrons,
which are in motion orbiting their atomic nuclei and spin-

0.1

0.2

0.3

time, milliseconds

0.4

0.5

ning on their own axis. Magnetic fields in electrical machines are created by electric currents flowing along conductors.
We will start with the fundamental phenomenon of force,
F, induced on a positive charge, q, in a conductive rod, by
moving the conductor with a constant velocity, v, at right
angles to a uniform magnetic field, B, Fig. A5.1. This force
is referred to as induced by the motion of the conductor
in the magnetic field. The force causes an accumulation of
positive charge at the top of the rod, as shown in Fig. A5.1.
Negative charges are driven in the opposite direction, and
accumulate at the bottom of the rod. The strength of the
electromotive force (EMF) is expressed in volts. A volt is
defined as the strength of electromotive force against which
a joule of the work is done, to move a coulomb of positive
charge. A coulomb of charge is defined as charge delivered
by a one ampere of current in one second.
We are accustomed to forces which are collinear with
motion, such as a force acting to accelerate a mass, m, per
FM = Ma , and the force on a charge, q, due to an electric
field E, per FE = q E. A force sideways to the direction of
motion seems odd at first, but we also deal with non-collinear forces in mechanics. Centrifugal and centripetal forces
act at right angles to the direction of rotational motion. A
challenge in developing a working understanding of electromagnetic phenomena lies in applying analyses previously
used in Dynamics and Fluid Mechanics to a new subject.
It is helpful for mechanical engineering students to keep
in mind that electromagnetic phenomena were discovered
and characterized by the forces associated with them. Our
primary interest in these phenomena is that we need those
forces, and the torques which can be created, to drive our
machines. A secondary benefit in studying these phenomena

Chapter 5 Appendix

313
magnetic field B

Force on Positive Charge q


F=qv B
Uniform magnetic
field B into the
plane of the paper

Conductor

+
+
+
+

thumB

Speed (velocity) of charge


Second finger

Positve Charge
induced Force

First finger

Velocity v

Fig. A5.2 Right-hand rule mnemonic for the cross product F = q v B

- Negative Charge

Fig. A5.1 Creation of electromotive force by the motion of a conductor


across a uniform magnetic field

is the strengthening of important analytical skills which have


wide application.
The cross product, F = q v B, defines the relative orientations of the force, the velocity of the charge, and the
magnetic field. A mnemonic is often helpful to visualize the
force induced on a charge during its motion through a magnetic field. The mnemonic relates the thumb and first (index)
and second (middle) fingers of the right hand, held such
that they form three orthogonal axes, to the three vectors in
F = q v B, Fig. A5.2. (This mnemonic allows engineering
students to make rude gestures and claim they were only
finding the direction of the induced electromotive force.) We
need to use speed as the synonym of velocity.
thumB=BFirst finger=Force
Second finger=Speed=velocity

Fig. A5.3 A Simple Generator


or Electric Motor

The small step from the fundamental physical phenomena,


illustrated in Fig. A5.1, to an electrical machine, a generator, or, the inverse, an electric motor, is to provide a path to
complete a circuit from the top to the bottom of the conductive rod moving in the magnetic field, Fig. A5.3.
Charge no longer accumulates at the ends of the moving conductor since the charge is now free to flow along the
stationary conductor. The EMF creates a current, i, which
flows through the stationary conductor from the positive end
of the moving conductor to the negative end. The strength,
or voltage v, of the EMF is proportional to both the velocity,
v, of the moving conductor and the strength of the magnetic
field, B.
Power is converted from mechanical power to electrical
power

P
= Fext v = iv
omitting the negative sign, which would be included using
linear graph sign convention.
If the externally applied force is the input to this system,
then this device is a generator converting mechanical power

Uniform magnetic field B into the plane of the paper


positive

Moving conductor
creates EMF and
current i

Externally applied force Fext

Velocity v

negative

Stationary conductor creates closed circuit

Current i

5 Fluid, Electrical, and Thermal Systems

314

i1 dL 1
L1

^
r

i1

permeability
of free space

12

Point A

L2

dFmagnetic = i 2 dL 2

i2

cross product
yielding the force
acting on i 2dL 2

Fig. A5.4 Two current carrying conductors. Current element i1dL1

to electrical power. If current is supplied by an external


source, then the device is a linear electric motor.

Electromagnetic Force between Two Current


Elements
The simple motor of Fig. A5.3 assumes the existence of a
uniform magnetic field. Although some electrical machines
contain permanent magnets, these are limited in size. The
motion of all large electrical machines is created by the electromagnetic force between two current carrying conductors,
either copper wires or bars. Both conductors create magnetic
fields. Although we will sometimes find it convenient to
perform calculations in terms of the interaction of the two
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
two cross products are required.
The electromagnetic force acting between two current-carrying conductors separated by air or in a vacuum, Fig. A5.4, is
expressed in the following differential vector equation.


d Fmagnetic =

4 r122

i2 d L 2 (i1d L1 r 12 )

cross product yielding


the magnetic field
acting on i 2 dL 2

(5.59)

where r 12 is the unit vector in the direction from element d L1


to d L 2.
Equation 5.59 is a differential equation, to allow us to address the general case, in which either wire 1 or wire 2 or
both are not straight. Infinitesimal wire segments, d L1 and
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.
Current is the flow rate charge. The product of current and
the length of a conductor are used, in place of charge measured in coulombs. The constant, 0, is the permeability of
free space,where free space is a physics term meaning a
volume free of material, i.e, a complete vacuum. The magnetic permeability of air, air , is approximately equal to that

0
_____
4 r122

(i dL r )
1

12

unit vector from

2 is the area dL1 to dL2


4r12
of a sphere with
radius r12

Fig. A5.5Annotation of the equation for the differential magnetic


force acting between two current elements

of a vacuum. The denominator term, 4 r122 , is the surface


area of a sphere with radius, r12 .
Although it is conventional to place scalar terms in front
of the cross products, the equation is easier to interpret when
it is rearranged as:
d Fmagnetic =


0
i2 d L 2 (i1d L1 r 12 )
4 r122

d Fmagnetic = i 2 d L 2

4 r122

(i d L

r12

(5.60)

and annotated as in Fig.A5.5.


What makes this equation difficult to visualize are the two
cross-products, Fig. A5.6. The cross-product, i1d L1 r12 , is
normal to the plane of i1d L1 and r12, when it is scaled by
the ratio,

0
, it is the magnetic field vector at the end of
4 r122

the position vector, r12. The density of the magnetic field


created by i1d L1 decreases as the inverse square of distance
from i1d L1. Consequently, the density of the magnetic field
created by i1d L1 is constant on any sphere centered at i1d L1.
The second cross-product, i2 d L 2

0
(i1d L1 r12 ), is
4 r122

the magnetic force acting on i2 d L 2. The vector, d Fmagnetic, is


normal to the plane of i2 d L 2 and i1d L1 r12.

The Magnetic Field B


The cross-product


dB =

0
r12
i1d L1
4 r122

(5.61)

Chapter 5 Appendix

315

i1 dL 1

sphere with
radius r12
centered at i1dL 1

s=0

i1dL1

12

i1dL1 r 12

i 2 dL 2 i1dL1 r 12

plane of cross-product

i1dL1 r 12

r12

i 2 dL 2 i1dL1 r 12

idL

i1dL1 r 12

is the infinitesimal portion of the magnetic field, dB, at the r12,


created by the i1d L1, and scaled by the permeability of free
space, 0 , and divided by the geometric factor, 4 r122 , equal
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 BiotSavart
law. It dates from 1820. It is a magnetostatic approximation
requiring steady currents. Time-varying currents, such as
AC, accelerate and decelerate charges, causing them to radiate electromagnetic energy. The magnetostatic approximation also neglects the dimensions and characteristics of the
conductors.
If we express r12 and r122 as functions of distance, s, along
wire, L1, then an evaluation of the resulting line integral over
the length of wire, L1, yields the magnetic field density, B, at
point A on wire L2,

0
i d L1 r12
i d L1 r12 ) = 0 1
2 (1
4
4 r12
r122

Be careful with the vector notation. The difference between


the last two expressions is easily overlooked. The vector r12
is the unit vector in the direction of r12, whereas the vector r12
is a position vector with magnitude, r12 = r12, Fig. A5.8.


r12 =

(5.62)

r12
r12

(5.63)

As one can easily imagine, integrals of this type range from


straightforward to very challenging, depending on the geometry of wire L1. We will deal with two simple geometries that
allow us to approximate all situations of interest to us. Wire
L1 will be either straight or bent into a circle.

Magnetic Field Density Calculations


Magnetic Field Density at a Distance x from a StraightWire-Carrying Current i
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 distance x from the wire in the normal plane passing through the
wires midpoint, at s = L0 /2.
L0

0 i1 d L1 r12
4
r123

Fig. A5.9 Magnetic field density at a distance x from a straight-wirecarrying current i

Fig. A5.7 The cross product i2 d L 2 (i1d L1 r12 )

B=

L0

Point A

i2dL 2

plane of cross-product

L0
__
2

i1dL1

i2

Fig. A5.8 Electromagnetic force acting at point A on current carrying


conductor L2 from infinitesimal current element, i1dL1

Fig. A5.6 The cross product i2 d L 2 (i1d L1 r12 )

L2

Point A

i2dL 2

B=

s=L

^
r

r12

i1

B=

0
i
(idL r ) = 0
4
4 r 2

L0

dL r
r2

d L r = d L r sin ( ) = d L sin ( )

5 Fluid, Electrical, and Thermal Systems

316
1.0

L0

i
2 L0
2
B = 0 2
2
2
2
4 x 4 x + L0 4 L0 s + 4 s

0.8

L0
________
4x 2 + L 02

0.6

x=1

0.4

x=5

0.2
0
0

10

L0

15

20

25

Fig. A5.10 Effect of the finite ends of a straight conductor on the


magnetic field density, B, on a plane normal to the midpoint of the
conductor. L0 is the length of the conductor. x is the distance from the
conductor

since
B=

r = 1.

0 i
4

L0

dL r 0 i
=
4
r2

dL r sin ( )

L0

r2

0 i
4

L0

dL sin ( )
r2

We will exploit the symmetry of the configuration and douL


ble the integral we evaluate, from s = 0 to s = 0 . We must
2
now express r 2 and sin ( ) in terms of x and s. The Pythagorean Theorem yields the length of the position vector
r squared
L

r = x + 0 s
2

sin ( ) =

B=

x
=
r

0 i

B straight =
wire

B straight =
wire

L0

s
x2 +
2

r122

0 i
4

1
L0

x +
s
2

L0

x2 +
s
2

L0

B=

0 i 2
2
4 0

2
L0

2
x
+

1.5

ds

ds

for

L0
<5
x

0 i
for
2 x

L0
5
x

L0
5 , the magnetic field density around a straight
x
current-carrying wire is inversely proportional to the distance, x, from the wire. Lines of equal strength of the field
are circles centered on the wire, in planes normal to the axis
of the wire. The sense of direction of the magnetic field is
In practice, the
established by the cross-product, id L r.
right-hand rule is used. One imagines grasping the wire with
ones right hand with ones thumb pointing in the direction of
the current. The fingers of the right hand then curl around the
wire in the direction of the magnetic field lines, Fig. A5.11.
We speak of a magnetic field as flowing, so that we
can use familiar physical phenomena to help us visualize the
physics of electromagnetism. The magnetic flux, (Greek
for f, as in flow), is the total magnetic field, which passes
through either half of a plane containing the wire. The SI unit
for magnetic flux is Wb, webers. The magnetic flux is a
When

d L sin ( )

0 i
L0
2 x 4 x 2 + L20

and

L0

0 i
L0
2 x 4 x 2 + L20

The effect of the ends of the wire on the density of the magnetic field at the center decreases, as the wire becomes longer. However, end effect is a function of both the length
of wire L0 and the distance x of point A from wire. A plot
L0
vs. L0, for values of x = 1 and
of the function,
4 x 2 + L20
x = 5 shows how rapidly it approaches a final value of unity,
Fig. A5.10. Whether one judges the end effect to be negligible depends on the precision needed in the analysis. A conL0
5, which corresponds
venient value for our purposes is
x
to an error of less than 2%.

L0

B=

B=

From the definition of sine and the Pythagorean Theorem,

L0
L0
2

L0

2
2
2

x 4 x 2 + L2 4 L L0 + 4 L0 x 4 x + L0

0
0
2
2

i
B = 0 2
4

Chapter 5 Appendix

317

Magnetic field line


Current

idL

Pt. A

r
2R

Fig. A5.11 Circular magnetic field line B around a straight wire carrying
current i

Magnetic field lines


Current

Fig. A5.13 Point A at the center of a loop wire carrying current i

Fig. A5.12 Magnetic field lines B, around a single loop of wire carrying current i

tributed by an infinitesimal element of the loop of wire, lies


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.
We will begin with the equation for magnetic field densityB
L0

scalar, and is the same across any surface that cuts across the
magnetic field, from the wire to infinity.
The amount of magnetic flux passing through a unit
area is the magnetic flux density, B, B = / area. The SI
unit for magnetic flux density B is T, teslas. Magnetic flux
density B is a vector. The variation of the magnitude and
direction magnetic flux density B with position (or in case of
a straight wire, with distance from the wire) is a vector field.
Rather than draw a multitude of tiny vectors to represent the
magnetic field, it is easier and more common to visualize
magnetic fields using field lines, also known by the older
term, lines of induction. The spacing of field lines is used
to convey the magnetic flux density. More tightly spaced
lines denote higher magnetic flux density. An important difference between electric fields and magnetic fields is that
magnetic field lines always form closed loops, which is the
basis for the flow analogy.
Magnetic Flux Density B at a Distance x within a Loop
of Wire Carrying Current i
There are two ways to strengthen the magnetic field created
by a given current. First, we can shape the wire. For a given
current, the magnetic field inside a loop of wire is substantially stronger than the magnetic field next to a straight wire,
Fig. A5.12.
We will consider two cases. First, we will calculate the
magnetic field density B at the center of the loop. Next, we
will formulate the general case to calculate the magnetic
field density B at any point within the loop.
Magnetic Flux Density B at the Center of the Loop
The radius of the wire loop is R, Fig. A5.13. The circumference of the loop is L0 = 2 R. The magnetic field, con-

B=

0
i
id L
r = 0
2
4
4 r

d L
r
2
r

L0

The distance to point A is constant and equal to the radius R,


since point A is at the center of the loop

B=

0 i d L
i
r
= 0 2
4 0 r 2
4 R
L0

L0

d L r
0

The cross-product is also constant, since the lengths of d L


r are constant, and the angle between them is constant
and
and equal to

= 90o .

( )

d L
r = dL
r sin 90o = d L

r = 1 and sin = 1. This leaves


since
2
B=

0 i
4

L0

i
d L r
= 0 2
4 R
r2

L0

dL
0

We will exploit the symmetry of this problem to simplify the


calculation, and express the integration in terms of the angle
, defined graphically above.
B center =
of loop

B center =
of loop

0 i
4 R 2

L0

dL=

0 i
4 R 2

R d
0

0 i
iR 2 i
2
R ] 0 = 0 2 = 0
2[
2R
4 R
4 R

5 Fluid, Electrical, and Thermal Systems

318

North

In electromagnetics, field lines exit the coil at the North pole


and enter the coil at the South pole. Within a coil, the magnetic 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
Field Lines B

Current i

 agnetic Moment and Engineering


M
Approximations of Magnetic Field Density
The solenoidal current density, j s , is a useful quantity in
many calculations.

South
Fig. A5.14 Cross-section of the magnetic field within and around a
coil

where the direction of the magnetic field is determined by


the right-hand rule and is out of the plane and normal to the
plane of the paper in this example.

Coils or Solenoids
A logical extension to a single loop of wire is to stack
up loops of wire, by creating a helical coil, Fig. A5.14.
The coil was invented by Ampere, who coined the term,
solenoid,Greek for channel, to describe it. Today, solenoid commonly means a linear electromechanical actuator.
The effect of strengthening the field by shaping the wire is
taken to its limit with toroidal coil, since the magnetic field
lines never leave the coil. A coil-wrapped toroidal core of
ferromagnetic material, discussed below, is known as a Hayward ring, after the American physicist from the late 1800s.

Magnetic Sign Convention


Historically, magnetic phenomena were first observed with
permanent magnets, called lode stones of the ore magnetite.
It was first recorded in China that permanent magnets could
be created by magnetizing iron needles with a lode stone. The
Chinese were also the first to use compasses. Well before electromagnets were invented, terminology from cartography had
been applied to the orientation of a magnet field of a compass needle within the magnetic field of the earth. (The earths
magnetic field is created by the motion of charge with the flow
of liquid iron in the earths core.)
Unfortunately, and oddly, the sign conventions for geomagnetism and electromagnetics are opposite to one another. In the geomagnetic sign convention, field lines enter the
earth at the North pole and exit the earth at the South pole.

js

N i mmf
=
l
l

(5.64)

The term, density, is odd in this context, since the denominator is length, not area. Its use derives from the model of an
ideal cylindrical conductor with zero wall thickness replacing the physical coil. The current Ni is distributed uniformly
along the length of the cylinder, and flows circumferentially
around the ideal conductor. The units of the solenoidal current density reflect recurring confusion due to angular displacement and rotation lacking units. The units are
Ni
amp ampturns
s
=
j = =

m
m
l
We will use the units, [ amp turns/m ], since they help remind us that the number of turns N is a factor.
A second derived quantity useful in describing permanent
magnets is the magnetic moment of a current-carrying
loop with radius R and cross-sectional area A
magnetic moment i R 2 = i A
Magnetic moment yields the torque exerted about the axis
of a loop with the current i when multiplied by the magnetic
field density B. If B is not normal to the plane of the current
loop, then the dot product between B and a normal to the
plane of the loop is used.
Although it is relatively easy to program magnetic field
calculations in MATLAB, or other languages, approximations based on the closed-form results that were available
prior to machine computation generally have sufficient accuracy for engineering. The discrepancy between the more
precise numerical results and the closed-form approximations will be far less significant, than the uncertainty and
error introduced by including ferromagnetic materials in the
flux path to increase the magnetic flux density.

Chapter 5 Appendix

319

The following approximations assume that the magnetic


flux density is uniform across the radius of a loop or a coil.
They are based on the closed-form results using the Biot
Sarvart model with infinite current density for a loop or coil
in air.

Single-Loop Approximations
Magnetic flux density within a loop of radius R
B=

0 i

2i R 2 0 2iA
B= 0
=
4 r 3
4 r 3
where A is cross-sectional area of the loop, and the product
iA is the magnetic moment of the loop. The magnetic moment is, indeed, the moment which would act on the loop to
align the loop with the magnetic field.

Coil Approximations
Magnetic flux density at the middle of a solenoid of sufficient length such that the end effects can be neglected

0 Ni
l

= 0 j s

Magnetic flux density at any radial location at the middle of


a solenoid of sufficient length such that the end effects can
be neglected
B=

0 Ni
2l

2.0

Magnetic
Field
Density
B
teslas

0 j s
2

Note that the magnetic flux density at the ends of a long coil
is half that at the middle.

Magnetic Permeability and Ferromagnetic


Materials
The ease with which a magnetic field can be created in a
volume or material is called the magnetic permeability, and
is given the symbol, (Greek for m). The field calculations
above assumed that the volume surrounding the currentcarrying wire was filled with air. The permeability of air is
approximately equal to that of a free space (a vacuum), 0.

saturation

1.5

preferred
operating
range

1.0
0.5

2R

Magnetic flux density outside a loop of radius R at a point


of distance r from the center of the loop, the far field approximation

B=

2.5

10

100

Applied Magnetic Field

100,000
10,000
.turns
amp
_____
Intensity, H
meter
1,000

Fig. A5.15 DC magnetization or BH curve

webers
henrys
= 4 10 7

meter
ampere turn meter

0 = 4 107

Although permeability is usually referred to as a material


property, this is not strictly correct, since a magnetic field
can permeate a vacuum in which there is no material. The
term, 4, is introduced for convenience when performing
calculations, since the magnetic density created by an infinitesimal current element obeys the inverse square law.
Ferromagnetic materials, iron, cobalt, nickel, and their alloys, have permeabilities significantly greater than that of
free space. We increase the strength of a magnetic field in
an electrical machine by creating a flux path for the magnetic field through ferromagnetic materials. Materials have
magnetic properties on the atomic level due to the motion of
their electrons. If the vectors which represent the axes magnetic dipoles created by the orbiting electron are randomly
oriented then there is no net magnetic field on the macroscopic scale. Ferromagnetic materials significantly enhance
the strength of an applied magnetic field because the magnetic dipoles created by orbiting electrons in neighboring
atoms align into magnetic domains. The magnetic dipoles
of the magnetic domains within the ferromagnetic material
orient in response to an externally applied field. However,
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.
The non-linearity of the relationship between the applied
external magnetic field strength and the resulting magnetic
field density can be seen in two plots used to characterize the
magnetic properties of materials. The first is a DC magnetization, or BH curve, which is a plot of the magnetic field
density which would be created in a ferromagnetic material
on its first exposure to an applied external magnetic field, Fig.
A5.15. The second non-linearity is the ferromagnetic materials memory, which is why a ferromagnetic coating is used

5 Fluid, Electrical, and Thermal Systems

320

for computer hard disks. The response of a material to its first


magnetization is a reproducible standard test, but doesnt reflect the behavior of the material in actual use, when it may
experience thousands or millions of cycles of magnetization.
There are three regions of interest in this S shape.
Note that it is a semi logarithmic plot. The initial region
has a low B / H slope, indicating that low applied magnetic field strength accomplishes disproportionately little
orientation of magnetic domains. Increasing H from 1 to 5
amp turns/meter increases B from approximately 0 to 0.25
teslas. The slope B / H then increases sharply, and it is
this region which is the preferred operating range, where we
can create a substantially more powerful magnetic field density with the least amount of current. Increasing H from 6 to
10 amp turns/meter increases B from approximately 0.3 to
1 teslas. As the slope of the curve lessens with increasing applied magnetic field strength, there is diminishing return for
the investment in electric current. Eventually, the maximum
possible orientation of the magnetic domains is achieved.
The magnetic field density will continue to increase with increasing applied magnetic field strength, but only at the rate
it would for air or a vacuum, since the ferromagnetic material can no longer amplify the magnetic field.
The magnetic properties of the ferromagnetic materials
vary widely between the different materials, between two
specimens of the same material processed differently, and
even in a specimen, as its magnetic history changes. Annealing an alloy resets the magnetic domains, by providing the
thermal energy necessary for them to return to their randomly oriented state.
The BH plot and plots of permeability and relative permeability for annealed iron are shown in Figs. A5.16 and
A5.17.
The BH curves for commercially available silicon steel
used in electrical machines often has two sets of units for
both axes. The units kilogauss and oersteds are cgs. The units
amp turns/meter and teslas are mks.
The magnetic properties of engineering ferromagnetic
materials alloys are dependent on their temperature history,
heat treating, or annealing. Just as one cannot have confidence in published strength values for steels, neither can one
have confidence in published magnetic properties, unless the
heat treatment of a specific part is known and identical to
that of the specimens the published data is based on. In fact,
magneto-inductive non-destructive testing is used to measure strength, surface hardness, and the depth of case hardening, since variation in these properties can be correlated with
variation in electrical conductivity and magnetic permeability. Conversely, manufacturing processes which affect these
mechanical properties also affect electrical conductivity and
magnetic permeability.
The second important plot of magnetic behavior is a hysteresis plot, Fig. A5.18. Hysteresis arises from the memory

3.5
3.0
2.5

Magnetic
Field 2.0
Density
1.5
B
teslas 1.0
0.5
0.0
10

100

1,000

10,000

100,000 1,000,000

Applied Magnetic Field Intensity, H

.turns
amp
_____
meter

Fig. A5.16 BH curve of annealed iron, data from Sears (1951)

of a ferromagnetic material. When the applied magnetic


field strength that has produced a state of magnetization is
reduced, magnetic domains lose orientation at a lower applied magnetic field strength than that at which they gained
orientation.
The area within the hysteresis loop is energy in joules per
cubic meter lost as heat to the ferromagnetic core. This can
be seen from the units of H and B
A Wb
A V sec
A V sec J
[H ][ B ] = 2 = 2 =
=
m m m m m3 m3
(5.65)
Hysteresis is important when the electrical power driving the
system is AC, since the energy loss due to hysteresis occurs
each cycle. The magnitude of the hysteresis energy loss reduces the energy efficiency of the device. The factor which
constrains the amount of electrical power that can be applied
to electrical machines is the temperature rise of the core and
the winding or coil. Energy lost as heat due to hysteresis of
the ferromagnetic core and electrical resistance of the winding both contribute to the temperature rise.
Because the electrical power which can be applied to the
machine is dependent on the internal temperature, many
mechanical engineers are engaged in heat transfer analyses
of electrical equipment. An example is the computer industry, where increasing density of integrated circuit chips has
led to more elaborate cooling schemes, including aluminum
heat sinks and fans mounted directly on central processing
units. Google, the internet search corporation, chose a site
directly adjacent to a river for its main computing facility,
because of the need to cool its machines.
Some magnetic domains will retain orientation at zero applied magnetic field strength, creating a permanent magnet, Fig. A5.19. A permanent magnet can be demagnetized
by reversing the externally applied magnetic field. The permanent magnets used in electrical machines were magne-

Chapter 5 Appendix

321

Fig. A5.17 Magnetic permeability and relative permeability


of annealed iron, data from Sears
(1951)

0.008

Permeability

amp
weber
______
meter

6,000
5,000

0.006

4,000
0.004

3,000

Relative
Permeability
r

2,000

0.002

1,000
0.000
10

100

1,000

10,000

0
100,000 1,000,000

Applied Magnetic Field Intensity, H

teslas

teslas

residual inductance
remanence or
retentivity

amp.turns
______
meter

Br

coercive force

Fig. A5.18 Magnetic hysteresis loop. H is the intensity of the applied


magnetic field. B is the magnetic flux density

tized carbon steel or cobalt steel, until the second half of the
twentieth century, when the harder, or more-difficult-todemagnetize ferromagnetic alloys Alnico (for aluminum,
nickel, and cobalt), were developed. Magnetic ceramic materials, which include rare-earth elements, were first introduced in the 1960s. Although they are mechanically brittle,
they have superior magnetic properties. Keep in mind even
the best permanent magnet can be demagnetized, by accidentally 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 permeability of a ferromagnetic material, where relative is with
regard to the permeability of free space. The relative permeability of a material, r, is defined as


turns
amp
_____
meter

(5.66)

Handbook and Web values of the relative permeability for


common engineering materials are reported as ranges, due to
the effects of processing on crystal structure.
Ferritesare ceramics composed of barium, manganese,
and/or zinc and iron oxide, with other materials as binder.
The relative permeability of ferrites varies widely, since
some are custom-formulated for different applications.

amp.turns

H ______
meter
Hc

Fig. A5.19 Portion of hysteresis loop showing the residual magnetic


flux density, Br, at zero applied magnetic field intensity, Hc

Table A5.1 Relative Permeability r


Iron
Nickel
Cobalt
Carbon Steels
Cast Irons
Stainless Steels
Electrical (silicon alloy) Steels

150 to 5,000
110 to 600
70 to 250
50 to 100
100 to 750
1 to 3,000
2,000 to 6,000

Low-cost barium-based ferrites have relative permeabilities


in the range of cast iron. Their advantage over cast iron is
their substantially lower electrical conductivity and, hence,
less heating due to eddy currents. Ferrites formulated to enhance magnetic fields under low frequency AC or DC excitation typically have relative permeabilities between 750
and 10,000, but some range up to 100,000. At the extremes,
magnetic glasses can be manufactured with relative permeabilities of 1,000,000. Ferrites can have both high magnetic
permeability and high electrical resistivity, as opposed to
electrical steel which has relatively low electrical resistivity.
Unwanted currents induced in electrical machinery by timevarying magnetic fields are called eddy currents. Eddy
currents are an energy loss and can be a significant problem in electrical machines because of the resulting resistive

5 Fluid, Electrical, and Thermal Systems

322

heating. The high electrical resistivity of ferrites reduces


eddy currents.

 agnetic Flux Density B, Flux , and Applied


M
Magnetic Field Intensity H
The relative permeability of the ferromagnetic materials in the magnetic flux path within an electrical machine
(ferrite, electrical steel, carbon steel, and cast iron) is large
enough that the portion of the magnetic flux outside of the
ferromagnetic material can be neglected. Ferromagnetic materials can be thought of as not only intensifying the magnetic field but also channeling it, since the magnetic field
outside of the ferromagnetic material is generally negligibly
small, compared to the field inside the ferromagnetic material. Hence, we will model the magnetic flux as confined
to, or channeled by, the ferromagnetic materials of the machine, with one important exception. Electric motors must
have gaps of some finite width between their stationary and
moving elements. The effect of gaps which are in the flux
path must be included in our analyses, as they are always
significant.
Magnetic flux density B is a vector field. Perhaps magnetic field would be a more straightforward, modern name,
but, as noted above, the lines of induction, or magnetic
field lines, form closed curves. As a result, our engineering
models are based on a flow analogy. The term, flux density,
reinforces the flow analogy.
As we calculated, the magnetic flux density is non-uniform across the radius of a coil with an air core. The magnetic flux density cannot be expected to be more uniform,
when the core is a ferromagnetic material with a non-linear
permeability. It is, in fact, generally a great deal less uniform. We will, nonetheless, make the additional engineering approximation that the magnetic flux density is uniform
across any surface through the flux path. A realistic model
of the magnetic field within an electrical machine requires a
finite-element model based on a well-characterized ferromagnetic material. However, as is the case in stress analysis,
premature use of finite-element models electromagnetics is
poor engineering, as they tend to lead to well analyzed, bad
designs. Far simpler models are always used for the initial
design. They greatly simplify calculations, while maintaining an acceptable accuracy to parse through design alternatives. Once a candidate design is identified, then a finiteelement model may be appropriate. For our purposes, the
simple models will suffice.
By confining the magnetic flux path to the machines
iron (ferromagnetic materials) and the gaps between the
stationary and moving elements, and assuming the flux density is uniform across the path, we can relate magnetic flux
and flux density B as:

B =

(5.67)

where A is the cross-sectional area of the magnetic flux path.


In order to calculate the magnetic flux density when the
flux path comprises materials with different permeabilities,
we must introduce the vector field, H, called the applied
magnetic field intensity or magnetic intensity. The magnetic flux density, B, is proportional to the vector field, H.
The proportionality constant is the magnetic permeability, .


B = H

(5.68)

Why work with B and H rather than just B? Why add the
complication of another vector field, if B and H are just
scalar multiples of each other? There are two reasons. First,
although we will make the engineering approximation that
there is a single value of permeability for a given material, as
shown above, the relationship between the applied magnetic
field intensity, H, and the resulting magnetic flux density,
B, is non-linear and multi-valued. The contribution of ferromagnetic materials to the magnetic flux density, B, eventually is also limited. The limit, called saturation,is reached
when all domains are oriented to the applied magnetic field.
Any further increase in the applied magnetic field intensity,
H, does increase the magnetic flux density, B, but the incremental increase is only at the permeability of free space, 0,
as if the ferromagnetic material were not present.
Second, some electrical machines of interest, specifically some stepping motors and the type of AC motor used in
some hybrid automobiles, use rotors which are permanent
magnets. The magnetic flux of rotating motors flows through
the rotor. The permanently magnetized rotor and the externally applied magnetic field together produce the magnetic
flux through the motor. The field, H, is used to account for
both effects.
Our analysis will exploit the similarities between the flow
of magnetic flux through an electrical machine, and the flow
of electric current in an electrical circuit comprising a voltage source and resistors. The analogy, called a magnetic
circuit, draws the parallel between (1) voltage (electromotive force) in an electric circuit and magnetomotive force in
a magnetic circuit; and (2) the resistance-to-current flow in
an electric circuit and amount of magnetomotive force needed to establish a magnetic flux in a magnetic circuit. The
drop in voltage in the direction of current flow in an electric circuit is proportional to the resistance of the elements
in the circuit. Similarly, in a magnetic circuit, the drop in
the magnetomotive force in the direction of the flow of
flux is proportional to the reluctance, R, of that portion
of the magnetic circuit. Reluctance R is a nineteenth-century
term, coined to draw the analogy between a purely resistive

Chapter 5 Appendix

323

Coil 2

Coil1

i
Surface defined by the
magnetic flux path
cutting the conductors

Magnetic flux path


assumed on centerline
Path length L c

Ferromagnetic core
Cross-sectional area A

Fig. A5.20 One wire wrapped in two coils in series on a ferromagnetic


core

electric circuit and the magnetic circuit model. The magnetic


circuit model is developed below.

Fig. A5.21 Ferromagnetic core with two coils cut by surface across
flux path

(counter-clockwise when looking down on a horizontal


cross-section). Consequently, current i creates mmf in the
same direction through both coils, Fig. A5.21. Hence, the
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 will now simplify our calculation of the magnetic flux
density produced by coils, by creating a lumped parameter
model. We will create a pseudo-circuit and, to make it useful,
we will simplify the geometry of the flux path and linearize
the magnetic phenomenon.
Coils are used in all motors to intensify the magnetic field
created by a current. The number of turns of wire in the coil,
N, times the current flowing through the wire, i, is known
by the nineteenth-century term magnetomotive force, abbreviated as mmf.


N2 turns

N1 turns

Coil 2

Coil1

N2 turns

N1 turns

mmf = N i

(5.69)

We will model the magnetic field through motors using what


is known as a magnetic circuit,in which the mmf drives the
flow of the magnetic field in analogy to a voltage source,
also known as electromotive force, driving the flow of electric current.
Although we will see that the ferromagnetic materials
used in electrical machines have non-linear magnetic properties, we will use the engineering approximation of superposition, and add the individual mmf of multiple current-carrying
coils to find the net mmf acting to drive the magnetic field.
A ferromagnetic core with one wire wrapped in two
coils in series is shown in Fig.A5.20. The direction of the
mmf produced by each core is determined by the right-hand
rule. (Imagine grasping the wire with ones thumb pointing
in the direction of the current. The magnetic field circles the
wire in the direction of ones fingers.) The same current, i,
passes both coils, which are wrapped in the same sense

We increase the current density without increasing the


amount of current, by increasing the number of turns in the
coil.
We will begin with the case, in which the mmf in the circuit is created by coils, and the flux path is through a ferromagnetic core with constant cross-section. Our second case
will include a gap in the flux path. The third case will include
a permanent magnet.
H is a vector field in the direction of the magnetic flux
density, B, but of a different, and much smaller, magnitude.
Excluding the presence of a permanent magnet in the flux
path, the magnitude of magnetic field intensity, H, is the
drop in the magnetomotive force, mmf, in the direction of
the flow of flux , divided by the length of the segment flux
path, l , where l is a vector pointing in the direction of the
flux .


H=

mmf
l

(5.70)

Magnetic field intensity H is the intensity of the effort required at a location along a flux path to maintain a magnetic
flux . For a given drop in magnetomotive force mmf along
path of length l , the magnetic flux increases with increasing magnetic permeability and increasing cross-sectional
area A of the flux path. The dependence of the magnetic flux
on a material property and the cross-section area A
allow us to build analogies between the creation of a magnetic flux and potential driven flows of fluid, heat, or electricity.

5 Fluid, Electrical, and Thermal Systems

324
Table A5.2 Basis of magnetic
circuit analogy. l is the length of
the flow or flux path and A is its
cross-sectional area.

Fluid
Volume flow rate, Q
Pressure, p
Fluid permeability, k

Electrical
Current, i
Voltage, v (emf)
Electrical conductivity,

Thermal
Heat flow rate, q
Temperature, T
Thermal conductivity, k

Fluid resistance, R f

Electrical resistance, R

Thermal resistance, RT

Flow equation

p = Rf Q

emf v = iR

q=

Flow resistance parameter relationship

Rf =

Flow variable
Driving potential
Flow-related material
property
Flow resistance
Elemental parameter

l
kA

The corresponding flow equation which relates the magnetomotive force to the resulting magnetic flux uses the
resistance-like magnetic reluctance, R


mmf = R

R=

l
A

Case I Magnetic Circuit with a Uniform Ferromagnetic


Core
A ferromagnetic core has two coils connected in series, as
shown in Fig.A5.20. Express the magnetic flux as a function of current i. We will use the magnetic circuit model:

mmf = mmf1 + mmf 2


Be careful with the direction of the mmf. The assigned positive direction for the magnetic flux is shown on the figure
as clockwise. Magnetomotive force which acts in that direction is positive. Magnetomotive force which acts in the opposite direction is negative. Use the right-thumb rule to
establish the direction of the mmf. Imagine grasping a turn
in Coil 1 with the thumb of your right hand pointing in the
direction of the current. If, on the side of the wire adjacent to
the ferromagnetic core, your fingers curl around the wire in
the clockwise direction, then the mmf is positive, otherwise

RT =

l
kA

mmf = N1i1 + N 2 i2
Since the same current flows through both Coil 1 and Coil 2,
i1 = i2 = i
Hence,
mmf = N1i + N 2 i = ( N1 + N 2 ) i
We now use the definition of magnetic reluctance R to establish its value.
R =

mmf = R
We will first establish the relationship between current i
and the mmf in the magnetic circuit. How do we handle
two coils on the same flux path? We use superposition. The
total, or net, mmf acting to establish (or drive, if you prefer)
the magnetic flux in the circuit is the sum of the mmf of the
two coils

RT

it is negative. In this magnetic circuit, Coils 1 and 2 both


contribute positive mmf. The total mmf is

(5.71)

where
Magnetic Reluctance
Flux PathLength
l
R =
=
Permeability Cross Sectional Area A

lc

Ac

Substituting into the magnetic circuit equation


l

( N1 + N 2 ) i = cA

and rearranging
Ac
i=
lc

( N1 + N 2 )

Example Magnetic Circuit


Consider the toroidal ferrite core and coil shown in
Fig. A5.22. A coil of N = 200 turns of wire, carrying a current, i = 3A, creates an electromagnet with a magnetomotive
force, mmf = N i . The length of the flux path, called the core
length, lc , is approximated as the circumference of the circle
with the mean radius of the torus, r = 0.5 in . There is a radial gap through the torus with gap length, g = 0.020 in. The
torus has cross-sectional area Ac = 0.0625 in 2. The relative

Chapter 5 Appendix

325

gap length g = 0.020 in

Core Reluctance
Torrodial ferrite core
Relative permeability r
Core length l c
Cross-sectional area A

Rc

R
mmf

Gap
Reluctance

Rg

3
200 turns of wire in the coil

Fig. A5.23
Magnetic circuit model. The magnetomotive force,
mmf=Ni=200 turns times 3 amps=600A-turns

Fig. A5.22 Ferrite torus with a gap of length g = 0.020 in . and a 200turn coil

permeability of the ferrite is r = 15, 000. Calculate the mag-

netic flux and the drop in mmf across the gap.


We will model this system as a magnetic circuit Fig. A5.23.
We assume that all of the magnetic flux is confined to the
ferrite core, except where the flux path crosses the air gap.
The mmf force is analogous to a voltage source. The magnetic reluctances of the ferrite torus and the air gap are in series
and are analogous to series electrical resistances.
Continuity

Node 1 = R c Node 2 R c = R g

Compatibility

mmf13 = mmf12 + mmf 23

Elements

mmf12 = R c R c mmf 23 = R g R g

Energy

No energy storage elements.

Before we can perform a reduction, we must first calculate


the magnetic reluctances of the ferrite torus and the air gap.
The calculation of the magnetic reluctances is analogous to
the above calculation of the electrical resistance of a solid
copper wire. The material property, magnetic permeability, takes the place of electrical conductivity. The geometric
quantities of cross-sectional area and flux-path length have
the same influence on the magnitude of the magnetic flux, as
they do on the magnitude of the current in the analogous resistive circuit. Increasing the cross-sectional area of the flux
path increases the flux pushed by a given mmf. Increasing
the length of the flux path decreases the flux. Calculation of
the magnetic reluctance, R, and the electrical resistance, R,
use relationships of the same functional form.
Electrical Resistance R =
and

Length

Area

Magnetic Reluctance R =

Length

Area

The magnetic reluctances of the core, R c, and the gap, R g,


are in series. They sum to create the reluctance of the magnetic circuit, R .
R = R c +R g
Electrical and magnetic calculations are performed using SI
(or, sometimes cgs) units. We need to convert the dimensions
given in US customary unit into SI.
The mean radius of the torus is 0.5in. The circumference
of the flux path, including the air gap, is
Circumference = D = 2 r

1m
= 0.080 m
Circumference = 2 0.5 in
39.37
in

The length of the air gap in meters is


1m
lg = 0.020 in
= 0.00051 m
39.37 in
The core length is the length of the flux path minus the gap
length
lc = 0.080 m 0.00051 m = 0.07949 m
The cross-sectional area of the core is assumed to equal the
cross-sectional area of the flux path through the gap. The
cross-sectional area in square meters is
6.452 10 4 m 2
4
2
Ac = 0.0625 in 2
= 4.03 10 m
1 in 2

The relative permeability of the ferrite is r = 15, 000. Recall that relative permeability is defined as:

5 Fluid, Electrical, and Thermal Systems

326

where the permeability of free space is 0 = 4 10 7 henry/m.


The symbol for henry is L. Take care that the unit for
inductance does not accidently morph into alength. Likewise, take care that the unit for current, A, does not morph
into an areaA.
We now calculate both reluctances using the relationship,
Length
.
R =
Area
R g calculation. Recall that the magnetic permeability of
air is approximately equal to that of free space.
Rc =
Rc =

Length
Length
=
Area 0 r Area

0.07949 m
4 10 7 henries
4
2

(15, 000) 4.03 10 m


m

R c = 1.05 105

mmf13 = R c + R g

A
weber

mmf13 = R c c + R g g

mmf13
Ni
=
Rc + R g Rc + R g

600 A turns
A
A
+ 1.00 107
1.05 10
weber
weber
5

600 A turns
= 5.9 10 5 webers
A
7
1.01 10
weber

Turns, the number of turns of the coil, is dimensionless.


It is common for coils in motors, solenoids, and transforms
to have many hundred turns of fine wire, so as to create a
substantial mmf with a small current.
Reduction 2 Input mmf13 = 600 amp turns, Output
mmf 23
mmf 23 = R g R g
mmf13
mmf 23 = R g

Rc + R g

This result has the same form as the calculation of the voltage across a resistor in a voltage divider.
A

1.00 107

weber
mmf 23 =
600 A turns
A
A
+ 1.00 107
1.01 105

weber
weber
A

1.00 107

weber 600 A turns = 594 A turns


mmf 23 =
A
1.01 107

weber

Note that the magnetic reluctance of the gap is 100 times


larger than the magnetic reluctance of the core.
Reduction 1 Input mmf13 = 600 amp turns, Output
mmf13 = mmf12 + mmf 23

Rg
mmf 23 =
Ni
Rc + R g

mmf 23 = R g
Rg
mmf 23 =
mmf13
Rc + R g

It is important to recognize that 99% of the applied mmf


is expended in creating the magnetic field across the gap.
Reducing the length of gaps in electrical machines greatly
increases the amount of electrical power they can convert to
mechanical power.

Calculation of the Magnetic Field Created


by the Poles of Permanent Magnet
Permanent magnets are used in stepper motors, DC servomotors, and, more recently, in AC motors used by hybrid
automobiles. A permanent magnet in the flux path increases
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
equivalent surface current, which is a fictitious coil with
the current that would create the same magnetization in the
volume, occupied by the permanent magnet.
We will assume that the strength of a permanent magnet
is known. However, in practice, the strength of a permanent
magnet is rarely known and must be measured. Briefly, the
standard technique for measuring the strength of a magnetic
field is to observe the dynamic response of a test coil first
placed in and then quickly removed from the magnetic field.
The test coil used is known as search, flip, or snatch
coil. The terms flip and snatch coils are descriptive of
the test process. In the case of a flip coil, a dynamic response of the test coil is created, by flipping or rotating the
orientation of the test coil from normal to the field lines to
parallel to the field lines, thereby reducing the flux linkage
quickly to zero. Similarly, a snatch coil is quickly moved,
or snatched, from the unknown field to a location where
the field is known. An alternate technique is to measure the
force exerted by the permanent magnet, which, as we will

Chapter 5 Appendix

327

see when we discuss electric motors, is directly related to the


magnetization.
There are three equivalent ways of expressing the strength
of a permanent magnet. As noted above, a permanent magnet
can be expressed in terms of the equivalent surface current,
which would replicate the magnetic field density, B. Alternatively, a permanent magnet can be described using the quantity, magnetization M, which is the magnetic moment per
unit volume. Lastly, the strength of a permanent magnet can
be described in terms of the strength of the magnetic poles.
The magnetization M of a cylindrical permanent magnet
with length l and cross-sectional area A is defined in terms of
the equivalent surface current as


Magnetization M =

( Ni )s
l

B=

Ni

lC

(5.73)

Pole Strength m
= M Magnetization
Area
A

rN

Point A
__

__

rS

m rN
________
HN =
__ 3
40 |rN |

__

- m rS
H S = ________
__ 3
40 |rS |

South
Fig. A5.24 Magnetic field intensity components at Point A due to
point magnetic poles

of its size relative to the distance to a location of interest, is


calculated as


(5.74)

Magnetization M is a vector quantity. Its direction is determined by the normal to the area, if the cross-sectional area
is a plane, and whether the magnetic pole is North or South.
If the cross-sectional area is not planar, then trigonometry,
geometry, or vector Calculus would be needed to determine
the direction of the magnetization vector, M. We will approximate the surface of the pole as a plane to simplify our
calculations.
The magnetic intensity, H, in a region surrounding a magnetic pole which can be considered a point pole, because

H=

1
40

m ri

i=0

(5.75)

where the position vector, r, is directed from the pole to the


point of interest, when m is a North pole and directed from
the point of interest to the pole, when m is a South pole. Note
that pole strength m scales the position vector r, i.e.,


Hence, magnetization M is the magnetic flux density of the


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 South magnetic poles are where the field lines (or lines
of induction) exit and enter the surface of a body, respectively. The spacing of field lines is a method of graphically
conveying flux density. The pole strength, m, is indicated
by the number of field lines. Conversely, the pole strength
per unit area is


__

(5.72)

where the subscript, s, denotes the equivalent surface current (or fictitious coil).
Note the equivalent surface current is assumed to magnetize a volume with the permeability of free space, 0. It
is instructive to compare this expression for magnetization
with the expression for the magnetic flux density, B, in a core
of length lc with permeability, .


North

Hi =

m ri

40 ri

(5.76)

and ri point in the same direction. There is no vector product


in this calculation.
Magnetic intensity H is a linear vector quantity. If the current flowing through a coil is doubled, then the magnitude of
the vector field representing the magnetic intensity is doubled. It is important to note the magnetic flux density vector
field is non-linear, in general, due to the non-linear magnetic
permeability of ferromagnetic materials.

Magnetic Circuit Calculation Including a Permanent


Magnet in the Flux Path
Salient poles on a motors stator (the stationary electromagnetic components fixed to a motors frame) are coils
with ferromagnetic cores, which extend toward the rotor, so
as to minimize the gap width and concentrate the magnetic
flux, Fig A5.25. Salient poles are found in some conventional electric motors and all stepper, or stepping, motors. A
stepper motor rotates from one equilibrium position to the
next, as the sets of coils on its salient poles are energized
with currents by a logic circuit, following an algorithm based
on the type of stepper motor, permanent magnet, hybrid, or

5 Fluid, Electrical, and Thermal Systems

328

using these values, the fictitious solenoid current, and the


input current.
Step 1 Determine the mmf of the permanent magnet. In the
abstract, the magnets strength is expressed in terms of pole
strength, m

Stator, ferromagnetic, permeability


Coil, N turns

i
Rotor
permanent magnet,
pole strength = m

l1
N
S

l1

l2

I.D. O.D.
g

Coil, N turns
w

Cross-section depth d into plane of paper

Pole Strength m
= M Magnetization
Area
A
We estimate the pole area, neglecting the curvature of the
rotor, as

Fig. A5.25 An electromagnetic model of one pair of poles of a permanent magnet or hybrid stepper motor

variable (switched) reluctance, and its use, high torque, high


speed, or fine position control. Briefly, stepper motors differ
by
the number of sets of coils,
whether there are two coils on one salient pole,
how many poles are spaced around the circumference of
the stator,
the number of stacks of poles in the stators axial direction,
whether the rotor is a permanent magnet, and, if it is, the
arrangement of the magnetic domains, and
whether the rotor is castellated (grooved to form small
salients).
All stepper motors rotate to reduce the reluctance of the flux
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
permanently magnetized rotors fall into two types, on the
basis of how the rotor is magnetized. The rotor of a permanent magnet stepper motor has magnetic domains which
reverse polarity 48 times in the circumferential direction,
Fig. A5.32. These motors are the least expensive type. The
rotor of hybrid stepper motor is polarized in the axial direction, Fig. A5.33. The term, hybrid, refers to the castellation of the rotor, which increases the variation in reluctance
hence the torque produced by the motor.
A schematic of one pair of poles of a permanent magnet
stepper motor is shown in Fig.A5.26. We will express the
magnetic flux through the rotor as a function of current i.
The analysis can be broken into three steps. First, we must
determine the mmf created by the permanent magnet. The
magnets mmf is then summed (superimposed) with the magnetomotive 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 magnetic core to calculate the flux through the system. Third, we
calculate the reluctances from the geometry and magnetic
permeabilities of the motor, and express the magnetic flux

Apole = wd
Hence, the magnetization, M, is
M=

m
m
=
A wd

and it is directed out of the North pole of the permanent magnet.


In practice, we cannot easily measure the pole strength
of a permanent magnet stepper motor directly. Instead, we
work backward from the fact that the magnetization, M, of
a permanent magnet is equal to its flux density, B, which we
calculate from the magnetic flux :
M=B=

We determine the magnetic flux from the relationship between the rate of change of the flux linkage, , and the voltage induced on the motors coil, when the permanent magnet
rotor is spun at a known angular frequency

v1g =

d
d
v1g = N
dt
dt

We are given the permeability of the permanent magnet as


m. Therefore, we can calculate the magnetic field intensity,
H, which the equivalent coil must provide to replace the permanent magnet in the magnetic circuit.
m
= M = B = m H equiv
wd
Recall the applied magnetic field intensity, H, is the mmf
with drops in the direction of the magnetic flux over a length
l in a material with permeability,

Chapter 5 Appendix

329

4
3

is

Ns

Fig. A5.26 Magnetic flux paths of the electromagnetic stepper motor


model

is

2
1

6
Fig. A5.27 Permanent magnet of the rotor replaced by the equivalent
solenoid

H=

mmf
l

B=

mmf
l

mmf =

lB

Substituting in the permeability, length, width, depth, and


pole strength of the permanent magnetic yields
mmf mag = ( Ni )S =

l2 m

mmf mag = ( Ni )S N S iS = 2 NiS =

m wd

where the subscript, S, on the equivalent mmfrepresents


the fictitious solenoid, Fig. A5.27.
The magnetic flux density in the permanent magnet, when
it is the magnetic circuit, is not equal to magnetization M! The
magnetization, M, is the flux density, when the flux path of
the magnet is through air. The flux density in the permanent
magnet, when it is in the magnetic circuit, will be higher, even
when there is no current through the coils, because the flux
from the magnet flows through a ferromagnetic core of high
magnetic permeability, rather than through air. We must calculate the flux through the magnet in the next step.
Step 2 Derive the equivalent magnetic circuit using the
mmf of the two physical coils and the fictitious coil to drive
the magnetic flux, . First, identify the magnetic flux paths.
Next, identify mmf nodes where the magnetic flux divides
or joins; or, where there is a change in the magnetic permeability or cross-section of the flux path. (We will add mmf
nodes to account for the sources of mmf, after we combine
the coils, to simplify the magnetic circuit.)
We will now simplify the magnetic circuit further, by determining the single source of mmf equivalent to the three
coils. The physical coils which carry the same current i can
be summed
mmf1 + mmf 2 = Ni + Ni = 2 Ni
To include the fictitious solenoid representing the permanent
magnet in the single equivalent mmf, we must let NS equal
twice N
NS = 2N

We can now establish the fictitious solenoids current using


our previous result.

iS =

l2

m
m w d

1 l2 m
2 N m wd

The single equivalent mmf is:

1 l2 m
mmf equiv = 2 N (i + iS ) = 2 N i +
2 N m wd
The equivalent mmf source will be placed in series with the
magnetic reluctances anywhere along the central path of the
magnetic circuit. An mmf node must be added to the circuit
to accommodate the mmf source element. Let us add node
7 between nodes 6 and 1, resulting in the magnetic circuit
shown in Fig. A5.28.
Due to the symmetry of the motor, all of the reluctances,
except that of the permanent magnet, R m , appear twice. The
reluctances of the poles and the gaps are in series. The reluctances of the two sides of the stator are in parallel. We will
reduce this circuit graphically, by combining reluctances
in series and reluctances in parallel, until there is just one
equivalent reluctance and mmf source in the circuit. Recall
that reluctances can be combined with the same rules as with
resistors. Reluctances in series sum
R pole + R gap + R m + R gap + R pole R equiv

series

2R pole + 2R gap + R m R equiv

series

Reluctances in parallel sum as inverses:


1
1
2
1
+
=

Rstator Rstator Rstator R equiv

parallel

R equiv

parallel

Rstator
2

5 Fluid, Electrical, and Thermal Systems

330

R pole

mmf

R stator

R gap

Rm

R pole

2R pole+2R gap+R m+

R gap

R
stator
_____
6

mmf

R stator
Fig. A5.30 Final magnetic circuit model of the motor with a single
equivalent magnetic reluctance

Fig. A5.28 Magnetic circuit model of the permanent magnetic stepper


motor

Flux lines exit


North pole

2R pole+2R gap+R m
5

mmf

R
stator
_____
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.
The final simplification is to sum the two equivalent reluctances in series to yield the circuit shown in Fig.A5.30.
Step 3 Calculate the reluctances from the geometry of the
motor and the magnetic permeability of the materials.
We will approximate the cross-sectional area of the gap as
a plane. Note the cylindrical surfaces of the gap faces, which
are necessary to allow rotation, also increase the surface
area. In reality, there is some magnetic flux which crossed
the gap from the sides of the ferromagnetic core near the
gap, rather than the face of the gap, further increasing the effective cross-sectional area of the gap. This phenomenon is
called fringing, Fig. A5.31. We will neglect it.
Reluctance calculations:

Fig. A5.32a Rotor of a permanent magnet stepper motor showing circumferential magnetic
domains, b Permanent magnet
stepper motor showing rotor and
stator with claw poles

South

North

Flux lines enter


South pole

Fringing

Length
g
=
Area 0 w d
R magnet =

lstator =

l1
Length
=
Area C w d

l2
Length
=
Area m w d

O.D. + I .D.
2

R stator

N S

S N

R pole =

The flux path through the annular stator is assumed to be at


the mean circumference.

Magnetized
Rotor Shell

(O.D. + I .D.)

2
4
lc
Length
=
=
Area C w d

Non-Ferromagnetic
Rotor Core or Yoke

Rg =

Polarity of the
Magnetic Domains

b
N

S
N

References and Suggested Reading

331

Fig. A5.33a Rotor of a hybrid


stepper motor with axial magnetic polarization showing castellation, b Hybrid stepper motor
showing rotor and stator poles
without their coils

b
North

Permanent
Magnet

South

We can now express the flux through the motor as a function


of motor current i.
mmf
mmf = R =
R

2 N (i + is )
2R pole + 2R gap + R m +

R stator
2

1 l2 m
2N i +
2
N m w d

(O.D. + I .D.)

l2
l1

C w d
g
2
2
+
+
+
w d w d
2
c w d
0
m

References and Suggested Reading


Bird RB, Steward WE, Lightfoot EN (2006) Transport phenomena,
2ndedn. Wiley, New York
Fitzgerald AE, Higginbotham DE, Grabel A (1981) Basic electrical
engineering. McGraw-Hill, New York
Fitzgerald AE, Kingsley C, Umans SD (2003) Electric machinery,
6thedn. McGraw-Hill, New York
Ogata K (2003) System dynamics, 4thedn. Prentice-Hall, Englewood
Cliffs
Rosenhow WM, Cho YI (1961) Heat, mass, and momentum transfer.
Prentice-Hall, Englewood Cliffs
Rosenhow WM, Hartnett JP, Cho YI (1998) Handbook of heat transfer,
3rdedn. McGraw Hill
Rowell D, Wormley DN (1997) System dynamics: an introduction.
Prentice-Hall, Upper Saddle River
Sears FW (1951) Electricity and magnetism, 2ndedn. Addison-Wesley,
Reading
Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
dynamics. Addison-Wesley, Reading

Power Transmission, Transformation,


and Conversion

Abstract

A simple machine, such as a lever, transforms the magnitudes of the forces and velocities 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.

6.1Introduction to Power Transmission,


Transformation, and Conversion
Two important classes of machine components are those
which (1) transmit and transform power and those which (2)
convert power from one form to another. Gear trains, chain
and belt drives, and levers and linkages transmit and transform power. Electric motors and solenoids convert electrical
power to mechanical power. Similarly, hydraulic motors and
actuators convert fluid power to mechanical power. Mechanical power is converted into electrical and fluid power by electrical generators and pumps.
The terminology used in system dynamics for these
classes of machine components are transformers and
transducers. Transformers and transducers are interfaces
which transmit power between subsystems in dynamic system models. They are the dynamic system elements that permit us to model useful systems, since most machines are a
combination of interacting subsystems. A transformer is a
machine element which links or interfaces two subsystems
of the same type of energy. A transducer interfaces subsystems of dissimilar types. Note that the terms transformer
and transducer have specific definitions in system dynamics
which differ from, but are based on, their common engineering usage.

It is essential to keep in mind that the energetic models


we create represent only the dominant energetic attributes
of a real system. We must keep our models simple enough
to be useful. All physically real machine components which
transmit, transform, or convert power have the energetic attributes of energy storage and dissipation, in addition to the
attribute of transmission, transformation, or conversion of
power. If the physical component also stores or dissipates a
significant amount of energy, then storage and dissipation elements must also be added to the model. Similarly, view the
power transmission, transformation, or conversion attribute
of a physical component as a distinct element of the model,
independent of the components other attributes.

6.1.1Transformers
In common usage, the term transformer refers to an electrical transformer which changes the magnitudes of voltage
and current in alternating current (AC) electrical power. Two
electrical subsystems are created, because electrical current
on one side of a transformer does not flow into the circuit
on the other side. There is a physical coupling between the
two subsystems but no direct electrical connection. The transformer 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

6 Power Transmission, Transformation, and Conversion

334

common electrical transformer is a step-down transformer


which decreases voltage and increases current, allowing AC
electrical power to be distributed at high voltage and low current to reduce resistive energy dissipation but used at safer,
lower voltages. We use electrical transformers as the basis of
our system dynamics terminology and define a transformer
to be the interface between subsystems of the same type of
energy. Be aware that in most engineering communication
a transformer is an electrical transformer. It is only in a
Systems Dynamics context that transformer has a broader
meaning.
As it happens, the physical coupling between the subsystems in an electrical transformer is magnetic and requires a
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
current in the primary circuit. Consequently, electrical transformers transform the sinusoidal current of AC power but
do not transform direct current (DC) power. Other devices
such as DC to DC converters or a DC motor driving a DC
generator must be used. We will defer our discussion of AC
power, until we have investigated the responses of systems
to sinusoidal inputs, known as their frequency response.

Pin = Pout

LA

Power
LB

F(t)
g

b
g

Subsystem I
Fig. 6.1 Levers and linkages interface translational mechanical subsystems

Power
Pinion
N1 teeth
Fluid Clutch
Damping b1

Bearing
Damping b 2

(t)
Inertia J
Angular velocity source.

Gear
N2 teeth

Subsystem II

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
energy conservation as equal power flows into and out of a
transformer


Pivot

Subsystem I

6.1.2 Ideal Transformers

Subsystem II

Lever

(6.1)

Magnitudes are used in Eq.6.1 to defer dealing with the sign


convention for the time being.
Naturally, real machine components which function as
transformers are made of real materials and have less than
ideal bearings or electrical conductors. Consequently, the
energetic model of machine elements which function as
transformers may need energy storage and dissipation elements, in addition to the element, which represents energy
transformation.
Examples of mechanical transformers include the following:
Levers and linkages which interface translational mechanical subsystems, Fig. 6.1
Gear sets which interface rotational mechanical subsystems, Fig. 6.2
Belt drives which interface rotational mechanical subsystems, Fig. 6.3

Double-ended pistons which interface fluid subsystems,


Fig. 6.4
The subsystems linked by transformers are comprised of energetic elements which handle the same type of power. For
example, the subsystems identified in the system with a lever
above are both translational mechanical subsystems Fig.6.5.
Likewise, subsystems in a system with the gear set and the
system with the belt drive are rotational. Transformers couple subsystems of the same energy domain or type (i.e.,
translational, rotational, fluid, or electrical domain).
Although the subsystems are linked by a transformer, they
are also separated by it. A transformer is an interface between
the subsystems which transmits power from one subsystem
to another. In an ideal transformer, all of the power which
leaves one subsystem is transmitted to the other subsystem
without energy storage or loss. The power is transformed
twice during the energy transfer, from its original form to
an intermediate form and then back to its original form. In
the translational system with the lever, power is applied to
the lever at node one, transmitted to node two as rotational
power, and then applied to the element acting at node two as
translational power, Fig.6.6.
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


Fig. 6.3 Belt drives interface
rotational mechanical subsystems

335
Compliant Shaft
Spring Constant K

Pulley
Diameter D1

Flywheel
Rotational Inertia J1
Flywheel
Rotational Inertia J2

Pulleys Diameters
D2 and D3

Subsystem I

Subsystem II

Angular Velocity
Input (t)

Power

Subsystem III

Power

Fluid Coupling
Damping b

Pulley
Diameter D4

Fluid Accumulator
Doubled-ended Capacitance C
massless piston
Area A1 Area A 2

Power
= FBv2g

Lever

Pivot

Resistance R

p(t)
Pump

Subsystem II

Subsystem I

Intermediate
Rotational
Subsystem

Power

Fig. 6.4 Double-ended pistons interface fluid subsystems

= FAv1g
Power

LB
LA

F(t)
g

Fluid
Reservoir

Subsystem II
Translation

b
g

Subsystem I
Translation

Fig. 6.6 Transformers rely on a second energy domain. A lever transfers translational power via rotational power

+x,v

FA
1

L1

FB
2

L2

+,
Fig. 6.5 Lever annotated to show the positive directions for translational and rotational velocity and displacement

which separates the two subsystems from each other. The


double conversion of power also permits a transformer to
change the relative magnitudes of the power variables. The
magnitude of the power flow, i.e., the product of the power
variables, is unchanged if the transformer is ideal.

Actual transformers always dissipate some energy as heat


due to friction, electrical resistance, or magnetic hysteresis
and store some amount of energy. As always, the question we
must address when deciding how to model a real transformer
is How much energy? where a significant amount or a
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 different energetic element in the model. For example, since a real
lever has compliance and friction in addition to its transformation 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
two branches, one for each subsystem, separated by a gap so
that the branches connect to different nodes. A dashed ellipse
links the branches on either side to indicate the physical coupling which transmits power across the gap.

6 Power Transmission, Transformation, and Conversion

336

LB
v1g = - ___
v2g
1

LA

FA

PowerA

FB

g
Fig. 6.7 The linear graph symbol for a transformer consists of two
branches, one in each subsystem, linked by a dashed ellipse which denotes the power transfer across the interface

6.1.3Transducers
In mechanical engineering, the term transducer most commonly refers to a sensor, such as a load cell, which emits an
electrical signal in response to a non-electrical input, such as
a force, in the case of a load cell. Sensors typically interface
two dissimilar energy systems. In system dynamics terminology, transducers interface subsystems of different types
of energy. There is a significant difference between a sensor
and a system dynamics transducer. Sensors produce signals,
which are information, not power. An ideal sensors signal
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,
the mechanical power needed to produce a force signal from
a load cell is extremely small, since a load cell actually senses elastic displacements, which, given the size of a load cell,
are very small. Although the use of term transducer as the
classification of machine elements which interface dissimilar types of energetic subsystems is logical, it is also unfortunate, because it causes misunderstanding between engineers.
It is best to use the term transducer only outside of system
dynamics to mean sensor to avoid misunderstanding.

6.1.4 Ideal Transducers


System dynamic transducers are similar to transformers in
that they interface the power flow between two subsystems.
They differ from transformers, because they interface dissimilar subsystems. Examples of transducers include these:
DC motors interface electrical and rotational mechanical
subsystems
Hydraulic pistons interface fluid and translational
mechanical subsystems
Racks and pinions interface translational and rotational
mechanical subsystems

Transformer
or
Transducer

Power B

Fig. 6.8 A block model showing the positive power flows of a transformer or transducer. Note that both power flows are positive into the
transducer or transformer

Pumps interface fluid and rotational or translational


mechanical subsystems
Real transducers have dynamic properties, in addition to the
ability to transduce power. Similarly, a real motor or generator stores kinetic energy in its rotational inertia, as well as
magnetic energy in the magnetic field created by the current
through its windings. It also dissipates energy as friction in
its bearings and electrical resistance its contacts and conductors. In a dynamic model, each significant energetic property
must be represented by a different element. If the mass of
a real lever is significant to the dynamic performance of a
system, then a mass element and a transformer element must
both be included in the lumped parameter dynamic model to
represent both the energy storage and transformation properties of the lever. If the rotational inertia and friction is significant in a real electric motor, then an inertia element, a
damping element, and a transducer element must be included
to represent these three independent energetic properties.

6.1.5 Block Model of Power Flows


The transformation or transduction of power across an interface between subsystems can be represented schematically
as a rectangle, or block, with the power flows on either
side of the interface both shown in the positive direction,
which is defined as into the interface, Fig.6.8. We will refer
to this as the bank account sign convention for power flow.
Power flow in is positive. Power flow out is negative.
Because a transformer or transducer neither stores nor dissipates energy, power which flows in one side must flow out
of the other. Consequently, although both flows are shown in
the positive direction as flowing into the element, this cannot
be. One power flow must be out. In the ideal case, power can
flow in either direction across the interface. Either PA or PB
can flow in and the other out at a given instant. Showing both
flows into the transformer or transducer underlines this fact.
Unfortunately, it also requires a negative sign to reverse the
flow direction of one of the two power flows



PA + PB = 0

(6.2)

PA = PB

(6.3)

6.1 Introduction to Power Transmission, Transformation, and Conversion

337

Assuming both flows are into the interface is a useful, general


purpose assumption, particularly for systems in which power
can oscillate back and forth across the interface. The negative
sign introduced so that power flows sum to zero works nicely
with gear sets, levers, belt drives, and other transformers. It
is less than helpful for motors, pumps, and other transducers,
where signs are a problem, due to incompatible sign conventions in the two subsystems. Ultimately, we must return to
the linear graph and the schematic model of the system to interpret the meaning of a sign. We will not be above fixing,
or, more politely, redefining, signs when necessary. A system
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.
However, for the time being, we will accept whichever sign
the linear graph system yields.

Fig. 6.9 Linear graph symbol


of the lever as a transformer.
Note that the arrowheads on both
branches of the symbol orient the
positive direction of the through
variables, FA and FB towards
ground. Also note that the
transformer is identified by one
of the two transformer equations,
which relate a power variable on
one side of the interface to the
corresponding power variable on
the other side

6.1.6 Transformer and Transducer Equations

v2 g = v1g
(6.4)

Transformer and transducer equations are derived, beginning


with energy conservation expressed in terms of power flows.

PA + PB = 0
This expression represents the power flow into the two ports
of the interface, but it does not provide the equations we
need to derive a differential system equation. We have two
power variables on each side of the transformer or transducer
interface for a total of four variables. Say we were to work
with an electrical transformer where

PA = FA v1g

and PB = FB v2 g

Expressing energy conservation in terms of power yields


FA v1g + FB v2 g = 0
This equation can be rearranged to express one variable in
terms of the other three, say
FA v1g
FB =

v2 g
Although it is a valid equation, is not a particularly useful
addition to our equation list, because we derive a system
equation by substituting to eliminate unwanted power variables. This equation would introduce three variables in place
of one.
We prefer equations which contain only two variables, so
that we do not add variables to the equation upon substi-

L
LA

B
v1g = - ___
v2g

FA

FB

tution. Consequently, instead of a single equation with four


variables, we will derive two equations formulated in terms
of a power variable from each side of the interface of the
form

and


FB =

FA

(6.5)

where is the transformer constant in this case, since the


variables on either side of the equation are from the same
type of energetic system, translational mechanical. We will
include transformer and transducer equations with the element equations in our list of equations extracted from the
linear graph. Notice that they have a form similar to a dissipative element, i.e., power variable = constant power
variable.
Derivation of the transformer constant can be done via
simple engineering analysis. On the other hand, test data are
usually needed to determine a transducer constant. We will
introduce the techniques required with the relevant transducer
or transformer.

6.1.7Transformer and Transducers Linear


Graph Symbol
The linear graph symbol for transformer and transducers
consists of two branches, one for each side of the interface,
Fig.6.9. The through variables on each side of the interface
must be named, such as FA and FB, so that they can be used
in equations. The transformer or transducer is identified by
an equation which relates a power variable on the left side
of the interface to a power variable on the right side, written
above the transformer or transducer symbol.
A dashed ellipse encircling the two branches of the interface indicates that there is a power flow across interface.

6 Power Transmission, Transformation, and Conversion

338

The through variables do not act or flow across the interface! There is no branch connecting nodes 1 and 2. This is
(or should be) obvious in a transducer, because the through
variables on either side of the interface are different. The
transducer element of a model of a DC motor separates the
electrical and rotational mechanical subsystems. The respective through variables are current and torque. There cannot
be a branch connecting the nodes across the interface in a
transducer, because the continuity equation summing currents and torques would not make sense. They are different
physical quantities. Although it is less obvious in a transformer, because the subsystems on either side have the same
units, a branch across the interface is just as meaningless.
FA and FB are different forces, because they act at different
points on the lever. They are coupled due to moment equilibrium about the pivot, not a force balance which includes
a mystery force flowing between nodes 1 and 2. There is no
branch connecting nodes 1 and 2.
The power flowing in any branch of a linear graph is the
product of the through and across variables of that branch.
Referring to the levers linear graph transformer symbol
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
interface. Energy conservation for the transformer, expressed
in terms of power, because an ideal transformer neither stores
nor dissipates energy, is

PA + PB = 0

PA = PB

This equation indicates that one of the two flows is opposite the assumed positive direction, towards ground. For the
example lever, one of the forces, FA and FB, must act in the
negative direction and the other in the positive direction, because both velocities, v1g and v2g, always have the same sign,
since they must be in the same direction.
In this text, transformer and transducer branches will
often be connected directly to ground. However, there is no
need for them to connect to ground. It is perhaps more common in machine design that they do not. For example, many
motors are attached to compliant fittings which incorporate
some damping. The motor mounts in an automobile include a compliant elastomer (a rubbery polymer) to reduce
the vibration transmitted to the passenger compartment. A
motor mount is modeled as a spring and dashpot in parallel.

6.2Transformers
6.2.1 Mechanical Transformers Levers
We will assume that the rotation and resulting horizontal
displacement of the lever shown in Fig.6.10 is small, so
we do not have to consider motion in the vertical direction.

+x,v

FA
1

FB

L1

L2

Fig. 6.10 A lever is a translational mechanical transformer. Note that


the forces acting on the lever at nodes 1 and 2 are both shown acting in
the positive direction

A = FA v1g

Lever
transformer
interface

B= FB v2g

Fig. 6.11 Block model of power flow through the transformer. Note
that both power flows are positive into the transformer

Within this assumption, the lever is a translational mechanical transformer. Further, we will assume that the only property of the device is transformation. The lever does not store
or dissipate energy. For this to be reasonably true, the lever
would need low mass (small kinetic energy storage), high
stiffness (small strain energy storage), and low friction at the
pivot (small energy dissipation). Define an arbitrary positive
direction, as shown. Assume that there are two translational
velocity nodes on the lever, nodes 1 and 2, and show forces
FA and FB acting at nodes 1 and 2 in the assumed positive
direction.
There is a physical coupling between the power flow into
the lever at node 1 and the power flow into the lever at node
2. However, there is no force which acts directly from node 1
to node 2. Our linear graph symbol will have a gap between
these two nodes to represent the interface.
Energy conservation for the lever can be expressed in
terms of power, since no energy is stored or dissipated by
an ideal transformer. The sum of the power flowing into the
lever equals zero, Fig.6.11.

PA + PB = 0
The power flowing into the lever at node 1 and node 2 is

PA = FA v1g

and PB = FB v2 g

since

PA + PB = 0 PA = PB

6.2Transformers

339

Then,
FA v1g = FB v2 g
If the lever can be modeled as rigid, then the translational velocity of nodes 1 and 2, relative to ground, are related. The
translational velocities, v1g and v2g, are both functions of the
angular velocity of the lever, , and the distance of the node
from the pivot
v1g = L1 and v2 g = L2
We eliminate the angular velocity, to relate v1g and v2g
directly
v1g
L1

==

v2 g
L2

v1g =

L1
L2

v2 g

Having established a relationship between v1g and v2g from


the geometry of the lever, we can now substitute into the
equation of energy conservation, expressed in terms of
power to determine a relationship between FA and FB

PA = PB

L
FA v1g = FB 2 v1g
L1

L
L
FA v1g = FB 2 v1g FB = 1 FA
L2
L1

L
v2 g = 2 v1g
L1

(6.6)

L1
FB = FA
L2

(6.7)

Note that there is a fractional and sign inversion in the constant term between these two equations. The fractional inversion is a result of energy conservation, PA = PB . The
sign inversion is a result of the linear graph method convention to assume both PA and PB flow into the transformer,
PA + PB = 0 . The sign inversion between FA and FB tells us
that one of the forces is acting in the negative direction.
It is easier to see the fractional and sign inversion of the
transformer ratio, when the transformer ratio is renamed,
L
n 2
L1
v2 g = n v1g

and

1
FB = FA
n

L
FA L1 + FB L2 = 0 FB = 1 FA
L2
It is generally preferable to use geometric compatibility of
velocities, rather than equilibrium of forces or moments,
when possible, for two reasons. First, equilibrium equations are more prone to sign errors of action versus reaction. It is easier to see the direction of a velocity than a force
or moment. Second, it is often more difficult to formulate
the equilibrium statement than the geometric compatibility
statement. Our lever analysis is a good example. Note that
summation of the horizontal forces shown is not a statement
of equilibrium
FA + FB 0

The two transformer equations for this lever are




All transformers and transducer equations have a fractional


and sign inversion of the transformer or transducer constant,
when they are written with the power variable from one side
or branch of the interface on the same side in both equations.
In the transformer equations above, the power variables for
PB , v2g and FB, are on the left side of both equations.
Although it is recommended, we did not need to derive
the transformer equation using the relationship between the
velocities of nodes 1 and 2. If we had preferred, we could
have used a statement of moment equilibrium about the
pivot. Defining counterclockwise as positive, with the forces
acting in the positive x-direction, as shown in Fig. 6.10

because we did not include the horizontal reaction force at


the pivot. That is why we wrote a moment equilibrium statement about the pivot. It was obvious in this case that we were
missing the reaction force, although there are many mechanisms where the relevant equilibrium condition or all of the
reaction forces and moments may not be as obvious.

6.2.1.1Mechanical Transformers Belt and Chain


Drives
Belt drives and chain drives are used to both distribute and/
or alter rotational mechanical power. They allow multiple
shafts at different locations and in different orientations to be
driven by the same motive source, Fig.6.12. They also allow
the magnitude of angular velocity and torque to be changed
as needed, within the constraint that their product does not
exceed the power available.
6.2.1.2 Belt Drives
Belt drives are used to drive the water pump, alternator,
power steering pump, and air conditioning compressor in
automobiles. Automobile engine fans used to be driven by
belt (or directly from the crank (main) shaft. They are now
more commonly driven electrically, so that they can be
turned on and off as needed. An automobiles air condition-

340

6 Power Transmission, Transformation, and Conversion

Fig. 6.12 A nineteenth-century


belt drive with pulley shafts at
90

vsurface

2g

1g

vsurface
Fig. 6.13 A belt drive schematic showing the nodes and sense of direction of rotational velocity and the surface speed of the belt

er compressor is turned on and off as needed but, because it


is belt driven, an expensive, magnetically engaged clutch is
required. A number of different style belts are used in machine design, but there are only two basic types: (1) smooth
and (2) cogged or toothed belts. The cross-section of a
smooth belt may be planar, v shaped, or ribbed. A belt is
smooth if the cross-section is constant in the longitudinal
direction along the belt. Conversely, the cross-section of
toothed belts varies periodically in the longitudinal direction. Rectangularly shaped, transverse projections (teeth)
are spaced uniformly along the belt to engage pockets on
thee pulleys, so that the belt cannot slip. Cogged or toothed
belts are commonly called timing belts, since they are
used to synchronize the rotation of an automobiles cam
shaft(s) and crank shaft in place of chain drives or gear
sets. Belt drives have applications in computer printers and
scanners. Toothed belts have replaced chain drives in many
applications, including as the main drive of motorcycles.
They have also replaced wire rope to lift elevators.

6.2.1.3 Chain Drives


Although chain drives can be more expensive and require
more maintenance than belt drives, they can also carry heavier loads and have longer service lives. A current application
of a chain drive is as a component in the automatic transmission for a front wheel drive car. The automatic transmission of a front wheel drive car must be shorter than one for
a rear wheel drive car, because the engine and transmission
of a front wheel drive car is mounted transverse, facing
the passengers side, rather than lengthwise, facing front. A
chain drive is used to transfer the power from the torque
converter pump and turbine shafts to a second, parallel shaft
with the clutch plates and gear sets, essentially splitting the
transmission into two to shorten it.

6.2.1.4Belt and Chain Drive Transformer


Equations
The analyses of a chain drive and belt drive are identical.
Chains are engaged by sprockets and cannot slip without
breaking. Although smooth belts can slip against a pulley,
producing a squeal, they are not supposed to. Therefore,
one presumes that the belt, like a chain, does not slip against
either driving or being driven by a pulley. If that is true, then
the surface speed of the pulley equals the surface speed of
the belt, Fig.6.13.
We will assume that we have ideal massless belts and
chains which are perfectly flexible, but completely inextensible, meaning that they conform to the sprocket or pulley
but do not stretch. If they are massless, then they cannot store
kinetic energy. If they cannot stretch, then they cannot store
elastic energy. We will assume that their perfect flexibility
allows them to conform to the sprocket or pulley without
dissipating energy in deformation. Lastly, we will assume
that the chain or belt slips on and off the sprocket or pulley
without friction. Although belts and chains are designed to
minimize energy storage and dissipation, none of these assumptions is true. If a significant amount of energy is stored
or dissipated in any of these mechanisms, then elements
must be added to the system to account for that energy.
At any instant, one side of the belt between the pulleys
will be slack and the other side taut. The taut side of the
belt conducts the tensile force from the driving pulley to the
driven pulley needed to transmit power. The slack side has
enough tensile force to maintain static friction between the
pulleys and the belt, as well as to keep the belt from flapping.
The slack side of the belt does not transmit power. Recall
the engineering adage, You cant push a rope. Belt drives
transmit power by pulling against the pulleys with the net
tensile force, FTaut FSlack = FNet . The power transmitted is
PBelt = FNet vsurface.

6.2Transformers

341

Expressing energy conservation in terms of power,


Fig.6.14, the power PA flowing into or out of the belt drive
at node 1 is

A= TA 4g

and the power PB flowing into or out of the belt drive at node
2 is

PB = TB 2 g

(6.8)

and
(6.9)
vsurface = rB 2 g
where rA and rB are the radii of the pulleys. Equating the
surface speeds yields
rA 4 g = rB 5 g
and, hence, the relationship between the two angular velocities
r
(6.10)
2 g = A 1g
rB
In practice, pulleys are identified by their diameter, not by
their radius. The relationship between the angular velocities
can be expressed in terms of diameters, if convenient


2g =

DA
1g
DB

r
TA 4 g = TB A 4 g
rB


Although we could formulate the transformer equations for


a belt drive using moment equilibrium in terms of FNet, it is
less error prone to use geometric compatibility. The surface
speed of the pulleys can be expressed in terms of their angular velocities and radii
vsurface = rA 1g

B= TB 5g

Fig. 6.14 Block model of positive power flows into the belt drive
transformer interface

PA = TA 1g

Belt Drive
transformer
interface

(6.11)

The relationship between the torques acting on the two pulleys can now be calculated using energy conservation expressed as power, since the belt drives transform attribute
only transforms the power variables. If energy is stored in
the belt drive, then kinetic or strain energy storage elements
must be added to the model. Remember the linear graph sign
conventions require that the power flowing into and out of
the belt drive sum to zero. Equating the power flows leads to
a sign error, Fig.6.14

PA + PB = 0 PA = PB TA 4 g = TB 5 g

r
TA = A TB
rB

(6.12)

Note that we derived the transformer equations using the relationships between the geometry of the belt drive and the
velocities of the pulleys. We could have used the fact that
belt tension force is equal at the two pulleys to derive the
transformer equations, starting with pulley torque expressed
in terms of the belt tension force. However, equilibrium analyses are prone to sign errors. The sign convention of the linear graph method requires the unknown torques acting on a
body to be positive (i.e., in the positive direction). Engineers
have the tendency to assign forces and torques, so as to place
an object in equilibrium, which inadvertently reverses one of
the two torques.

6.2.2Gears Mechanical Transformers


and Transducers
Gear sets are used to either distribute or alter rotational mechanical power. They allow multiple shafts at different locations and
in different orientations to be driven by the same motive source.
They also allow the relationship between angular velocity and
torque to be changed as needed. Typically, gear sets step down,
or reduce angular velocity, and step up, or increase torque.
Overdrive refers to the event, when a gear set increases angular velocity above the velocity of the input shaft. For example, an
automobile may use overdrive at highway speed, when the transmission steps up the angular velocity of the drive (or propeller)
shaft, so that it exceeds the angular velocity of the crank shaft.
Gears are classified by the shape of the blank, the location of the teeth on the blank, the cross-sectional shape of the
teeth, the line or curve a tooth makes on the gear blank, and
the orientation of the gears shafts.
The overall shape of a gear is generally either a cylinder,
the frustum of a right cone (a truncated cone), or a rectangular prism (or bar), Fig.6.15. More exotic shapes are used
in some highly engineered applications, such as automotive
transmissions and differentials.
Gears can be combined in a wide variety of ways, since
the only fundamental requirement is that the teeth mesh as
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

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

Fig. 6.17 Bevel (left) and miter (right) gears. The axes of rotation of
miter gears intersect at right angles. If the angle of intersection is not
90, then the gears are bevel gears

The left-hand gear set in Fig.6.16 is a set of external spur


gears. The larger gear of the right-hand set is a ring gear.
Ring gears can have teeth on their cylindrical or planar surfaces.
Bevel and miter blanks are frustums of right cones,
Fig.6.17. Varying the included angle at the tip of the cone
allows the gears shaft to intersect at various angles. Miter
gears shafts intersect at 90o.
A rack and pinion, Fig. 6.18, is a transducer between rotational and translational motion, making it a transducer in
system dynamics terminology. A worm and gear has a high
gear ratio and is self-locking, meaning the gear cannot
drive the worm. The worm must drive the gear.

6.2.2.1 Spur Gears


Two intermeshing spur gears are shown in cross-section in
Fig. 6.19. The name spur refers to the similarity of the
cross-section to a horse riders spur. A spur gear has a cylindrically shaped blank. Its teeth have the involute shape,
which has the benefit of the two mating tooth surface rolling

against each other, rather than sliding, thus reducing the frictional losses and tooth wear. Consequently, excessive wear
on gear teeth indicates that the gears are either misaligned,
with their shafts incorrectly spaced or not parallel, or the
gears have different pressure angles, defined below, and
incompatible shapes.
A positive direction for the angular velocities and shaft
torques must be defined vectorially using the right-hand rule.
The pinion rotates at angular velocity 1g and the gear rotates at 2g. Torque TA acts on the pinion shaft and torque TB
on the gear shaft, both in the positive direction.
We can see from the schematic of the two gears, Fig.6.19,
that they must rotate in opposite directions (counter-rotate).
Both gears must have the same surface velocity of the pitch
surface, so that the teeth mesh. The smaller pinion must rotate faster than the larger gear to feed teeth into the nip at the
same rate. For gears to mesh, their teeth must be the same
size. Consequently, a tooth and the adjacent space can be
used as a convenient unit of measure, when comparing the
circumferences of two meshing gears. The formal measurements require identifying where on the surface of the tooth
the gear force acts. A circle passing through this location on
each tooth and centered on the axis of the gear is called the
pitch circle, Fig.6.20. Its radius and diameter are the pitch
radius and pitch diameter, respectively. The circular pitch is

6.2Transformers

343

Pitch Circle
Circular Pitch
N Teeth
Pitch Diameter
Fig. 6.20 The pitch circle, pitch diameter, and circumferential pitch
of a spur gear

Addendum
Dedendum

Pitch Circle

FT

Tooth Force FT

30 Helix

0 Helix

Fig. 6.22 External spur gears of two 0 and 30 helix angles

A= TA 1g

B = TB 2g

Gear Set
transducer
interface

Fig. 6.23 Block diagram of a gear sets positive power flows

Fig. 6.24 Linear graph symbol


of gear set

N
N1

2
1g = - ___
2g

FT
Pressure Angle
14-1/2 o, 20o, 25o

TA

Fig. 6.21 Tooth forces, pitch circles, and the addendum and dedendum

the circumferential pitch. The circular pitch is the distance


from the surface of one tooth to the next, along the pitch
circle.
The equilibrium condition for gears can be somewhat
counterintuitive. Levers have moment equilibrium about the
pivot. One can think of the teeth of meshed spur gears as a
series of levers making and breaking contact with one another. The contact force between the meshed gear teeth (the
tooth force) must be equal and opposite (action and reaction). The torques on the respective shafts are not equal and
opposite, unless the gears have the same pitch radius, where
the pitch radius is the distance from the axis of a gear to
where the tooth forces act.
The geometry of gears is described by the nineteenthcentury terminology, Fig.6.21. The addendum and dedendum are the length of the gear tooth above and below
the pitch circle. The pressure angle is the inclination of the
tooth force relative to the normal to the line between the two
gears centers of rotation.
The tops of spur gear teeth can be either parallel to
the axis of the gear or trace a helix about the gears axis,
Fig.6.22. The latter are known as helical spur gears. Mesh-

TB

ing gear teeth appear as mirror images of each other, when


looking at the nip, where they mesh. Helical gears carry
heavy torques, because more than one tooth is meshed. Helical gears also place a component of the tooth force parallel to
the gears shaft, which must be carried by the thrust component of the shafts bearings.
The relationships between the input and output torques
and angular velocities are calculated using energy conservation but expressed in terms of power flow, since the transformer attribute of the gear set neither stores nor dissipates
energy. If the actual gear set stores or dissipates a significant
amount of energy, where significant is always relative to
the other elements in the dynamic system, then elements
must be added to the model of the gear, to account for these
other power flows.
Proceeding as with the lever, power PA flowing into or
out of the pinion is PA = TA 1g and the power PA flowing
into or out of the gear is PB = TB 2 g , Figs.6.23 and 6.24.
The sum of the power flows into the transformer interface
is zero:

PA + PB = 0

PA = PB

TA 1g = TB 2 g

6 Power Transmission, Transformation, and Conversion

344
Fig. 6.25 60 Bevel gears

60 o
Shaft Angle

Pitch radius is an inconvenient and difficult dimension to


measure. It is far easier to count the number of teeth on a
gear, Nteeth, as a measure of its size of relative gears which
will mesh with it. The relative speeds of two gears pitch
surfaces must be equal for the teeth to mesh. The smaller
diameter gear must rotate faster to feed teeth into the nip at
the same rate as the larger gear


v pitch = rp 1g = rG 2 g
surface

(6.13)

Note the negative sign. Two meshing external gears must


counter-rotate (i.e., rotate in opposite directions).
The geometric relationship between radius, diameter, and
circumference are
radius =

diameter
2

and circumference = diameter

radius =

circumference
2

Using dtooth as the circumferential distance along the gear


from a point on a tooth to the same point on the adjacent
tooth, the circumference and radius of a gear is
circumference = dtooth N teeth radius =

dtooth N teeth
2

When two gears mesh, we can use the number of teeth on


each gear as a relative measure of its pitch radius


1g =

N2
2g
N1

(6.14)

Once the angular velocities of the two gears are related by


their respective teeth numbers, we substitute back into the
equation for energy conservation, expressed in terms of

power for the gear set to derive the relationship between TA


and TB
N
TA 1g = TB 2 g TA 2 2 g = TB 2 g
N1
N
(6.15)
TA = 1 TB
N2

6.2.2.2 Bevel Gears


The pitch surface of bevel gears is the frustum of a right
circular cone, Figs.6.25 and 6.26. They are used to change
the orientation of the shafts. If the shaft angle is 90o, then
the bevel gears are called miter gears.
6.2.2.3 Worm and Gear
Worm drives provide a large increase in torque and reduction in angular velocity, Fig.6.27. Rotation of the worm can
drive the mating gear in either direction. However, geometry
prevents rotation of the gear from driving the worm. A worm
drive is known as a self-locking gear set. As a result, a
worm drive is an extremely non-linear system, since power
can only flow in one direction, from the worm to the gear.
A good example of the importance of an understanding of
system dynamics in machine design is to consider the consequence of using a worm and gear to drive a large inertial
load. What could happen if the motor driving the worm were
to suddenly fail?
6.2.2.4 Gear Trains or Sets
Gear trains and gear sets are synonyms. Gear trains
transmit power between the input and output shafts with
some frictional loss in each nip (or mesh). Our analysis of
a pair of spur gears can be extended to any number of gears
arranged in a set. A particular gear set may be intended either
to (1) step up or step down (i.e., transform), the angular velocity (including direction) or the torque between the input

6.2Transformers

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
location or orientation of the output shaft relative to the input
shaft.
Determining the direction of rotation of the output shaft
is an elementary but error-prone step. Mechanical engineers
like to think that they can look at a gear set of any complexity, and see the relationship between the input and output
shafts rotations. This engineering hubris often yields needless, embarrassing errors. A reliable method of determining
the relative rotation of gears is to indicate the surface velocity at either side of every mesh, Fig.6.28.
The rotation reverses with each mesh from input 1g to
output 4 g . The relationship between angular velocities 1g
and 4 g is
N1 N 2 N 3
N N = 4 g
N

2
3
4

1g

1g

N1 N 2 N 3
= 4 g


N 2 N3 N 4

N1
1g = 4 g
N4

1g
N1

4g
N2
N3
N4

Fig. 6.28 The surface velocities are indicated on both sides of each
mesh to determine the directions of rotation

Notice that the presence of gears two and three does not
change the magnitude of the output angular velocity, but
their meshes do contribute to its direction. Since each mesh
reverses direction, an even count of meshes yields the direction of rotation for the input and output shafts. We can
confirm this by removing gear three from the gear train,
Fig.6.29.

N1 N 2

= 4 g
N 2 N 4

1g

N N2
1g 1
= 4 g
N2 N4

6 Power Transmission, Transformation, and Conversion

346

1g

N1

4g
N1

N2

N3
TA 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 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

PowerA
Driving Idler
Pinion

Flow In is Positive
Transformer
or
Transducer

Power B

Flow Out is Negative


Driven Gear

Fig. 6.30 Idler gears reverse rotation and change the location of the
output shaft but do not change the gear ratio of a gear set

N1
1g = 4 g
N4
Gears two and three in Fig.6.30 are called idler gears.
They serve to transmit power and change the location of
the output shaft but have no effect on the magnitudes of the
power variables. Idler gears do not affect the gear ratio of
a gear set. Again, idler gears do affect the direction of the
output shafts rotation.

6.2.2.5 Transforming Angular Velocity or Torque


As derived above, idler gears do not affect the magnitude
of the output torque or angular velocity, just their sign and
the location of the output shaft. The largest gear ratio which
can be achieved by a pair of gears is limited by geometric
compatibility and the space available. For a given tooth size,
the pinion has a minimum circumference that will allow its
teeth to mesh with the gear. The maximum allowable diameter of a gear is limited by the space available in the machine
and the machines budget. If the gear ratio needed cannot be
achieved with two gears, the alternative is to affix intermediate pairs of gears to common shafts. This arrangement is
known as a cluster gear or a compound gear. Each gear
on a common shaft has the same angular velocity, but its

Fig. 6.32 The power flows on both sides of the transformer or transducer are assumed to flow in the positive direction, into the transformer
or transducer. One flow must actually be in the opposite direction

surface speed varies with radius or number of teeth. Since


meshing gears have the same surface speed, this results in
the gear ratio of each mesh contributing to the inputoutput
gear ratio, Fig.6.31.
N1 N 3
N N 1g = 4 g
2

N1 N 3
1g = 4 g
N2 N4

6.3Transformer and Transducers Sign


Conventions
The linear graph method uses a sign convention for transducers and transformers, which presumes that power may flow
in either direction across the interface, Fig.6.32. The sign
convention has two parts. First, the assumed positive direction of the through variable in a branch of the transformer
or transducer is oriented in the direction of the drop of the
across variable of that branch, i.e., towards ground, Fig.6.33.

PA = TA 2 g and PB = TB 3 g
PA + PB = 0 PA = PB

6.3 Transformer and Transducers Sign Conventions

N
N2

3
2g = - ___
3g

PowerA

Power B
TB

TA

g
Fig. 6.33 The linear graph sign convention orients the power flows
towards ground down both of branches of a transformer or a transducer.
One power flow must actually be in the opposite direction. The dashed
ellipse indicates transmission of power across the interface

TA 2 g + TB 3 g = 0 TA 2 g = TB 3 g
The linear graph sign convention for transformers and transducers assumes that the power flows on both sides are positive, where positive power flow is defined as flow into the
interface. This is impossible, because an ideal transducer or
transform is an interface between subsystems which cannot
store energy. Hence, the second part of the sign convention
is that the two power flows must sum to zero. Physically, at
any instant, power flows into the interface on one side yields
power flow out of the interface on the other side. Reversal of
one of the two power flows inverts the sign of one of the four
power variables. Unfortunately, this sign inversion often
leads to nonsensical mismatches between the signs of different subsystems.
Signs are a problem in many mechanical engineering
calculations. Sign conventions used in the various subdisciplines within mechanical engineering arose organically
to suit the specific subdiscipline. The state of a variable that
was defined as positive was either (1) the same as a different
vector variable in the same system, (2) the more common
state, or (3) the preferred state. An example of the first case
is the definition in vector mechanics, that a positive force
accelerates a mass in the positive direction. It is possible to
invert the signs between displacement and force vectors, but
we would rather not. First, why bother? Further, sign inversions lead to errors. Negative signs are dropped or assigned
to the wrong variable. Therefore, we define a force accelerating a mass as positive, if the acceleration occurs in the
positive direction.
Examples of the positive direction as the more common
state arise, due to the nature of the materials in the subdiscipline. Particulates such as soil cannot carry tension. They
must be loaded in compression. Therefore, in soil mechanics, compressive force and compressive stress are positive.
Likewise, fluids cannot carry tension, except in carefully

347

contrived situations. Consequently, in fluid mechanics, positive fluid pressure is compressive. Negative pressure (tension) 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
reduces its bulk modulus to an unknown value. The bubbles
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 pressure. Cavitation can erode hardened steel.
When the positive direction is arbitrary, in the sense that
the variable is equally likely to have either state, engineers
tend to define the state we prefer as positive. For example,
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
the positive direction on one face, out of two normal to an
axis. Note that the sign convention places the element in tension. This jibes with the preference of mechanical designers,
although that may be a coincidence. Engineers prefer positive stress. Given two mechanical components of the same
length, the component loaded in tension would be smaller
in cross-section, requiring less material, than the component
loaded in compression. Further, tensile rupture is typically
preceded by some amount of plastic deformation, which can
signal the overload and forewarn of an impending failure.
Compressive failure is often catastrophic buckling, which
provides no early warning.
Unfortunately, solid mechanics and vector mechanics do
not have the same sign convention. Positive external forces
and internal forces are defined differently. A positive external force is a force which accelerates a mass in the positive
direction. A positive internal force yields tensile stress. What
is the state of stress of an element subjected to acceleration?
If the acceleration is large enough, so that the acceleration
of gravity is negligible, the element is compressed due to inertia, regardless of the direction of the acceleration. A given
force may produce positive acceleration (in the positive direction) and negative (compressive) stress in the same machine element.
As one would expect, signs become arbitrary, when we interface dissimilar subsystems. For example, what is the positive direction of rotation for the shaft of an electric motor?
There is some consensus that positive rotation is clockwise,
but there is no consensus where the observer stands, in front
of the motor or behind the motor. The only way to determine
the positive direction of a motor is to apply a positive voltage
(if it is a DC motor) and observe its rotation. The positive
direction of a motor depends on its manufacturer.
Our solution to the sign problems inherent in system dynamics will be to accept the signs of the linear graph method,
even when some are nonsensical. We will follow the linear

6 Power Transmission, Transformation, and Conversion

348

Pinion
N1 teeth

Inertia J

Fluid Coupling
Damping b

T(t)
Torque source T(t).
Positive rotation direction
defined by the right-hand rule

Gear
N 2 teeth

Fig. 6.34 Rotational mechanical system. The torque source defines the
positive direction for the system

graph methods sign convention during the reduction of a


system to a system equation. We will then define the positive
state for the output variable on a case-by-case basis, when
we interpret our results. Attempting to correct signs before
one has completely derived the system equation leads to errors and confusion. It is better to have nonsensical but consistent signs than uncertainty over what a sign means.
A common use of the linear graph method is to derive
the system equation (mathematical model) of a system to be
placed under closed-loop feedback control. In any feedback
control application, we must define signs such that a positive input yields a positive output. Otherwise, the resulting
closed-loop control system will be unstable.
In summary, the linear graph method sign convention for
mechanical transformers and transducers requires us to
1. Assume both power flows are positive, which is defined
as into the transformer or transducer.
2. The sum of power flows into the transformer or transducer must equal zero, because it cannot store energy,
PA + PB = 0. This inverts one of the power flow signs,
PA = PB .
3. Show the forces or torques acting on the transducer or
transformer as a free body diagram in the positive direction.
4. Orient the through variable flow on both branches of the
linear graph symbol towards ground.

6.3.1Example 1: The Linear Graph


for a Rotational System
with a Transformer.

The linear graph sign conventions are illustrated, using the


rotational mechanism shown in Fig.6.34, which consists of
a torque source, T(t), a fluid coupling or drag cup with rotational damping b, a spur gear set, and a mass moment of
inertia or rotational inertia J. The positive direction is defined as the direction of the input torque. Note that the shafts

in this system are not assigned energetic properties. This


implies that the shafts are very stiff, have negligible inertia,
and, consequently, can be modeled as ideal shafts having no
energetic properties. Ideal shafts have a single angular velocity, 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 velocity, . The torque source is shown as a vector. A real torque
source must react against something. Since no reaction
is shown, we assume that the torque source reacts against
ground. To transmit torque through the fluid coupling, the
drag and the cup (inner and outer elements) of the fluid
coupling must shear the fluid between them. Therefore, the
drag and cup and the shafts attached to them have different angular velocities. The output shaft of the fluid coupling
drives the pinion. The pinion and gear must counter-rotate.
Consequently, the pinion and gear will have different angular
velocities, even if they have the same number of teeth. The
inertia is fixed to the ideal shaft driven by the gear and has
the same velocity as the gear.
We have identified three distinct angular velocities plus
ground. The transformer relationship we derived above will
be used here, revising the subscripts for the angular velocities to be 2g for the pinion and 3g for the gear

2g =

N2
N
3 g and TA = 1 TB
N1
N2

Both of these equations are added to our equation list with


the other element equations. One of these equations must be
written above the transformer symbol on the linear graph to
identify the transformer. One equation is sufficient, because
the second transformer equation can be derived from it and
the linear graph. The two branches which form the interface
represent the power flows into the transformer:

PA + PB = 0 TA 2 g + TB 3 g = 0
N
N
TA 2 g TB 1 2 g = 0 TA = 1 TB
N2
N2
We need to define the positive direction for torques and angular velocities. If a positive direction is not indicated explicitly, then use the direction of the source. Because the
gear set is an interface between two subsystems, it has two
branches in its linear graph symbol. The torques shown on
the branches of the linear graph symbol for the transducer are the torques acting on the gears. The positive directions of these torques is towards ground. The linear graph is
Fig.6.35.
Now, we shall check that our sign conventions for the
transformer lead to signs that make sense. Torques TA and TB

6.3 Transformer and Transducers Sign Conventions

N
N1

349

2
2g = - ___
3g

Mass-moment
of Inertia J

TJ
T(t)

TP
g

TG

TB

Fig. 6.37 Torque Tb acting on the gear shaft and torque TJ acting to
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

Fluid Coupling
Damping b

Tb

TA
2

TJ
TB
Tb

TA

Fig. 6.36 Positive torques TA and TB shown on the shafts of the gear
set. Reaction torques Tb and TJ shown on nodes 2 and 3

are shown on the linear graph as positive torques, where the


positive direction for a through variable in the linear graph
is towards ground. Torques TA and TB are shown in the free
body diagram acting on shafts of the gear set, Fig.6.36. The
positive direction for this free body diagram was defined as
the direction of the input torque.
The reaction torque from the fluid coupling against the
pinions shaft, Tb, at node 2 and the reaction torque from
the rotational inertia against the gear, TJ, at node 3 are also
shown. Torques Tb and TJ shown are the torques acting on
the fluid coupling and the rotational inertia, respectively.
These are the torques, whose signs must make sense physically. Starting with the inertia, we see it will rotate in the
same direction as the gear, counterclockwise for the given
input torque. Using the right-hand rule on the vector TJ, we
see that it produces the correct counterclockwise rotation of
the inertia, Fig.6.37.
Considering the fluid coupling, Fig.6.38, we must have
equal and opposite torques acting on its two shafts. Torque TA
at node 2 is the torque acting on the pinion. It is in the positive direction. The reaction torque is the torque Tb, acting in
the negative direction. Similarly, at node 1, torque Tb is the
torque acting on the fluid coupling. The reaction torque T is
the torque of the fluid coupling on the torque source.

Tb
1

Fig. 6.38 Action and reaction torques shown on the fluid coupling

There are two aspects of this drawing which may look


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
in quasi-static equilibrium. (The prefix quasi means as
if or in a sense. In this context, it means we can formulate torque equilibrium, even though the fluid coupling is in
motion.) The torques acting on the fluid coupling must sum
to zero. The elemental equation for the fluid coupling with
damping b
Tb = b12 = b ( 1 2 )
states there must be a difference in angular velocity between
the two ends shafts of the coupling, for a torque to exist. The
two components of fluid coupling need not move in opposite
directions. In fact, in this system, they will rotate in the same
direction, expect during the transient period after the input
torque changes direction.
Second, if we have assumed that the applied torque, T,
acts away from ground, why is torque T at node 1 acting
towards ground? This is also a case of action and reaction.
If we were to draw a free body diagram of the device serving as our torque source, Fig.6.39, we would see that, for
the torque source to be in quasi-static equilibrium, and for
torque T to act away from ground at the ground (node g), the
torque T acting at the other end of the torque source at node1
must be in the opposite direction.

6 Power Transmission, Transformation, and Conversion

350

T(t)
Torque Source

Tb
1

Diametrically opposed graphite


brushes pressed against the
commutator by soft springs
energize the rotor windings

-i

Cross-section of
rolling-contact
bearing

Rotor, winds, shaft


Rotational inertia J

Tg

T(t)
g

Fig. 6.39 Action and reaction torques shown on the torque source

In summary, the linear graph sign conventions allow us


to yield the correct signs for the output variables of this system. This is true in the general case for a linear graph which
contains transformers but not transducers. Transformers interface subsystems of the same type. If we respect the linear graph sign conventions, the signs of the power variables
are consistent. We avoid the necessity of drawing free body
diagrams for every mechanical element in a system, saving
effort and the possibility of an action versus reaction error.
We will not enjoy the same success, when we work with
transducers. Incompatibility of the signs between systems of
different energy types requires further interpretation.

6.4Transducers
In systems dynamics, transduction is the conversion of
power between subsystems with different energy domains
or types. It is a reasonable term, but its meaning in system
dynamics does not correspond to its general engineering
usage, where a transducer is a sensor. Consequently, outside
systems dynamics, it is best to use the name of the type of
transducer, such as DC motor, or gear pump, rather than the
classification, transducer, to avoid confusion.
As an aside, there is weak analogy between the two uses of
transducer which is the root of its system dynamics usage. A
pressure transducer, for example, senses a fluid variable, and
produces a signal which is an electrical variable. Sensors interface between dissimilar systems. However, in the ideal, the
signals produced by sensors are not power flows but information flows. Ideally, the pressure sensors output voltage could
be read by the receiving instrument without a current flow into
the receiver. In reality, that never happens. All receivers draw
some power from the signal. Consequently, all signals contain
some power. Nevertheless, we model signals as being devoid
of power, except when we are designing the fan-out, the
number of receivers we can send a signal to, and have to deal
with the realities of our electronics.

Magnetic flux emerging


from the rotors north pole

+i

Commutator,
Rotating electrical
contacts

Energizing
current i

Fig. 6.40 DC motor rotor showing the commutator, brushes, and one
set of windings

6.4.1DC Electric Motors: Electrical


to Mechanical Transducers
Motor action results from the interaction of a motors magnetic fields. Large numbers of different designs lead to no
standard terminology. In general, motors consist of a stator
which is fixed to the motors frame, and, in the case of rotatory 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
magnetic field of the rotor is created by a set of windings
energized progressively as the rotor spins.

6.4.1.1 Permanent Magnet DC Motor


DC motors interface electrical and rotational mechanical
subsystems. A permanent magnet DC motor rotates, because
the magnetic field, created by passing a current through a
set of windings on its rotor, is not aligned with the magnetic
field of the permanent magnets adhered to the stationary
housing. The moment created by the misalignment rotates
the rotor to align the two magnetic fields. Before the fields
are aligned, the stationary contacts, the brushes, and the pair
of moving contacts on the shaft connected to the winding
break contact, de-energizing the winding. The stationary
brushes make contact with the next pair of contacts on the
shaft, as they move under them, energizing the next winding, which too is out of alignment with the stationary magnetic field, repeating the process. The ring of contacts on the
shaft is called the commutator, since it moves with the shaft.
The brushes are actually blocks of graphite, pressed against
the commutator by a spring to maintain contact. A drawing
of the rotor of a DC motor and schematic for a DC motor
model, neglecting the induction of the windings, are shown
in Figs.6.40 and 6.41.

6.4Transducers

351

A= iMv2g

R
v(t)

MOTOR

JM
bM

Fig. 6.41 DC motor energetic schematic. The motors induction is not


included in this model, because it occurs on a much smaller time scale
than the mechanical response of the motor

Other than the voltage source, all elements shown in this


schematic are energetic properties of the DC motor. The circle labeled motor represents only the transducer property
of the motor. The conversion of electrical power to mechanical power is known as motor action. There are other energetic properties in a DC motor, which are included in the
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 discontinuity between the graphite brushes and the copper commutator, known as contact resistance. These three electrical
resistances are in series. If the motor is poorly designed or
damped, arcing, an additional electrical resistance, can occur,
when electrical contact is broken between a pair of contacts
on the commutator and the brushes. Lastly, there is magnetic
hysteresis, which results in some of the energy stored in the
magnetic field not returning to the winding as electric current, but being lost as heat, when the coil is de-energized
and the magnetic field collapses. All of these energy losses
are summed and represented as a lumped property, an ideal
linear electrical resistance R.
The rotating element in the motor is made of a stack of
stampings of electrical steel (alloyed with silicon for high
magnetic permeability, to enhance the magnetic field created
by the current through the windings) supported by a steel
shaft. The surfaces of the steel stampings are insulated to
limit eddy currents created by the voltage, induced in the
steel by the movement of the rotor in the field of the permanent magnets. The windings are many turns of thin copper
wire. Copper is denser than steel, and the windings contribute substantially to the mass moment of inertia of the rotor,
J. In addition to kinetic energy storage in rotational inertia
J, there is frictional energy loss in the bushings, or bearings,
which support the shaft, as well as between the graphite
brushes and the commutator. Finally, the electrical steel of
the rotor becomes magnetized. There is magnetic interaction
between the rotor and stationary permanent magnets, when
the motor is de-energized. This phenomenon is known as
cogging, because of the variation in torque needed to ro-

B= TM 3g

DC Motor
transducer
interface

Fig. 6.42 Block diagram showing the positive power flows into the
DC motor transducer element, which interfaces the electrical and mechanical subsystems
Fig. 6.43 Linear graph
symbol of the DC motor
transducer interface

N
N1

2
1g = - ___
2g

TA

TB

tate the de-energized motor through a complete rotation. All


of these energy losses are lumped into a single ideal viscous
friction element, and modeled as rotational damping b.
The transducer property of the DC motor (its motor action), indicated in the schematic as a circle labeled motor,
is the interface between the electrical and mechanical subsystems. It is important to emphasize that the ideal transducer property neither stores nor dissipates energy. Any energy
storage or dissipation in a real transducer must be represented by elements in addition to the transducer element. The
transduction can be represented using the transformer block,
with both power flows flowing in, as shown in Fig.6.42, although one flow must be negative. In the linear graph symbol, both through variables, current iM and torque TM, flow
to ground, Fig.6.43. Both across variables, voltage v2g and
angular velocity 3g, are referenced to the ground.
The product of current iM and voltage between node 2 and
ground on the electrical side of the interface is the electrical
power, which flows into the transducer


PElectrical = iM v2 g

(6.16)

Likewise, the product of torque TM and angular velocity between node 3 and ground on the electrical side of the interface is the mechanical power, which flows into the transducer


PMechanical = TM 3 g

(6.17)

We must experimentally determine (or be provided) one of


the two transducer equations. Assume that we experimentally determined the relationship between the current flow

6 Power Transmission, Transformation, and Conversion

352

through the motor, and the torque produced by the motor to


be reasonably approximated by this linear relationship:


TM = KT iM

(6.18)

where KT is the motors torque constant.


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.
If we know the experimentally determined relationship,
TM = KT iM , then we can use conservation of energy, expressed in terms of power, to derive the relationship between
the voltage drop across the transducer, v2g, and the angular
velocity, 3g,

PA + PB = 0 PA = PB
iM v2 g = TM 3 g

iM v2 g = KT iM 3 g

v2 g = KT 3 g
(6.19)
The voltage, v2g, across the electrical branch of the transducer, is known as the motors back EMF, where EMF is the
abbreviation for electromotive force. Electromotive force is
a nineteenth-century term for magnetically induced voltage.
A conductor moving across a magnetic field, such that it cuts
the field lines (Chap.5 Appendix), creates a voltage in the
conductor. This is the principle of operation of electrical generators. The windings on the rotor of a motor cut through the
stators magnetic field lines create a voltage which opposes
the voltage from the source driving the motor. This principle
is Lenzs law, proposed by Heinrich Lenz in 1833. The symbol L for inductance is for Lenz. Induction acts to limit the
speed of electric motors. Even a frictionless electric motor
has a maximum speed, which is the angular velocity at which
the motors back EMF is equal in magnitude to the voltage
applied by the power source.
Checking the units of a system equation is more difficult,
if there is a transducer in the system, because the transducer
constant appears to have different units in each of the transducer equations. In fact, the two sets of units are the same, if
expressed in fundamental units. However, do not check the
units of the system equation, by expressing the transducer
constant in fundamental units. It is easier to simply check the
system equation units, by expressing the transducer constant
units in terms of both sets of power variable units, and then
making the substitution for the transducer constant units last,
when it will be clear which set to use, rather than by expressing all of the parameters of the system in fundamental units,
which often produces more errors than it reveals.
Example: Express the units of motor torque constant KT in
terms of the power variables

[T ] = [ KT i ]

[ K ] = i

[ v ] = [ KT ]

[ KT ] =

and
v

The terms gyrating and non-gyrating transducers are


definitions that academics use, which carry no practical significance. A non-gyrating transducer is one whose power
equations relate the same-type variables to one another.
The DC motor is non-gyrating, because it relates torque
to current, and these are both through variables. A gyrating
transducer relates a through variable in one subsystem to an
across variable in the other.
The sign conventions for electromechanical transducers
(motors) are arbitrary. Although one would expect a convention to exist, there is none. The direction of rotation of a motors shaft depends on position, relative to the front of the
motor. Consequently, how one looks at the motor must be
defined for clockwise or counterclockwise to have meaning.
The motion of a motor in response to an applied voltage is
consistent, relative to the housing of the motor, but independent of the orientation of the motor in space. Pick a motor
up and turn it around. It will rotate in the opposite direction,
relative to a fixed reference frame. Some motors are dual
shafted (which is really a single shaft that extends out the
front and the back of the housing), to further complicate
signs. The only useful definition of the positive direction of
a motors rotation is the direction it rotates, when the voltage
applied is positive.

6.4.1.2Solenoids
Ampere coined the term solenoid, Greek for channel, to describe the effect of multiple turns of a coil at intensifying the
strength of magnetic flux , created by a current i. The term,
solenoid, is now commonly understood to describe a linear
electric motor, which is actuated by a coil around a cylindrical ferromagnetic core, Fig.6.44. The block model of the solenoid and its linear graph are shown in Figs.6.45 and 6.46.
Another style of solenoid is shown in Fig.6.47. Energizing the coil moves the ferromagnetic plunger, thereby
maximizing magnetic flux by minimizing magnetic reluctance R. All motors can be understood from the perspective
of either motion, resulting from interacting magnetic forces
or torques or motion to maximize magnetic flux .
6.4.1.3Example 1: The Linear Graph of the DC
Motor Model
The first step in drawing a linear graph is to find the nodes
of distinct values of the across variables on the schematic,
Fig.6.48. What are the across variables of the two subsystems

6.4Transducers

353

FM

Normally-closed
contact

Spring contact

____
1CR

Logic Output

1CR

Normally-open
contact

Logic Input

R axial

shell

air gap

Coil 1CR

Fig. 6.44 Schematic of a solenoid in an electromechanical relay. Energizing 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

A= i Mv2g

R iron

Solenoid
transducer
interface

B = FM v3g

R iron

air gap

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, minimizing the magnetic reluctance of the magnetic circuit

v(t)

MOTOR

iM

JM
bM

FM = KF i M
2

2
3

Fig. 6.45 Block diagram of positive power flows into a solenoid linear
electrical motor transducer element. The symbol for voltage and translational velocity are the same. The node subscripts identify the element
that the variable is associated with, thus, distinguishing voltages and
velocities
Fig. 6.46 Linear graph symbol
of the solenoid (linear electric
motor) transducer interface

R radial

plunger

g
Fig. 6.48 DC motor model showing nodes of distinct values of voltage
and angular velocity

FM
1

TM = K T i M
2

g
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)
and the mechanical non-accelerating inertial reference velocity (also called ground) are two different physical quantities, they are both represented by the ground plane of the
schematic for graphical convenience.
Draw the linear graph, by first drawing and identifying
the nodes, and then adding the linear graph elements between
the corresponding nodes, Fig.6.49. There must be a transducer interface symbol in the linear graph to separate the two
subsystems. Remember, the linear graph symbol for a transducer has two branches which represent the two sides of the
interface. A branch in the linear graph can only represent one
energetic property. A common error is to use a branch in the
linear graph, which happens to be parallel to the interface, as

v(t)

iM
g

TM

Fig. 6.49 Linear graph of the DC motor model of Fig.6.48

one of the two branches. Do not assign a branch more than


one property by adding transduction to it. One way to avoid
making this error is to draw the interface first and assign
unique variable names to the through variables in the two
branches of the transducers linear graph symbol. Assigning
through variable names makes it less likely that a parameter
will also be mistakenly assigned to a transducer branch. Finally, remember to identify the transducer with one of its two
transducer equations.

6 Power Transmission, Transformation, and Conversion

354

R
v(t)

MOTOR

JM

JL

bM

v(t)

v(t)

MOTOR

JM

TM
g

Generator
Output
Terminal

Rwinding

JL
bL
g

Fig. 6.51 Schematic with nodes of distinct values of the across variables voltage and angular velocity

6.4.1.4Example 2: The Linear Graph for a DC


Motor Driving a Load through
a Compliant Shaft
A dynamic system with a permanent magnet DC motor driven by a voltage source is shown schematically in Fig.6.50.
The torque produced by the motor is proportional to the current flow through the motor, TM=KT iM. The motor has electrical resistance R, rotational inertia JM, and viscous friction
bM. The motor drives an inertial load, JL, through a shaft with
appreciable compliance, described by torsional stiffness K.
The inertial load rotates on hydrodynamic bearings with viscous friction bL.
Again, the first step in drawing a linear graph is to find
the nodes of the distinct values of the across variables, and
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
variable of the rotational mechanical system is angular velocity.
Draw the linear graph, Fig.6.52, by drawing and identifying the nodes, and then add the elements between them. Remember that the transducer property of the motor interfaces
the electrical and mechanical subsystems. The interface is
between the electrical resistor R and the rotational inertia JM.
One would extract the energetic equation from the linear
graph of this system as usual. This system has three independent energy storage elements and two rotational inertias
separated by the rotational spring. We will not reduce a system as complex as this to a single differential equation. We

JM bM

JL

bL

Fig. 6.52 Linear graph of the system shown schematically in Figs.6.50


and 6.51

bM

2
3

TM = K T i M

iM

bL

Fig. 6.50 Schematic of a DC motor driving an inertial load through a


compliant shaft

T(t)

J
b

DC
Generator

External
Electrical
Load

R Load 2
RLoad1 L

Fig. 6.53 Schematic of a DC generator driven by a torque source

will perform a partial reduction to a set of three first-order


differential equations using the state-space method.

6.4.2Generators: Mechanical to Electrical


Transducers
The back EMF of a DC motor is the voltage produced on the
rotors windings as they move through the stators magnetic
field. A DC motor functions as a DC generator, when the
motors shaft is driven by a mechanical power source with
the intent of converting the mechanical power into electrical
power, Figs.6.53 and 6.54.

6.4.3 Pumps: Mechanical to Fluid Transducers


Mechanical to fluid transducers are machines which convert
translational or rotational mechanical power to fluid power.
The system dynamics terminology is awkward, since fluid
systems are mechanical systems in that they obey Newtons
laws.
Pumps are classified by the type of fluid they move (either
gas or liquid); the motion of the mechanical element(s) which
moves the fluid; either rotational or reciprocating translational motion; and whether fluid is moved by trapping it in a
volume, or by establishing a pressure gradient. Pumps which
expel fluid trapped in a volume are positive displacement

6.4Transducers

355

Fig. 6.54 Linear graph of a DC


generator and electrical load
driven by a torque source

v2g = KG 1g

T(t) J b TG
g
pumps. Pumps which move fluid by establishing a pressure
differential in the fluid include fans, turbines, and diaphragm
pumps. Fluid can flow through a turbine pump, when the
pump is not driven. Conversely, fluid cannot flow through a
positive displacement pump unless it is driven.
The clash of sign conventions between the solid, vector,
and fluid mechanics is beyond awkward. It can lead to contradictory motions between solid elements and fluids. Unfortunately, the sign conventions used in solid mechanics and
in fluid mechanics are opposite. Positive stress in a solid is
tensile stress, whereas positive pressure compresses a fluid.
Negative (absolute) pressure places a fluid in tension. This
state of affairs is a result of sign conventions being established separately in each engineering specialty. The general
notion in each field is reasonable enough. Typically, if opposite states are possible, then the positive sign indicates what
an engineer prefers. If only one state is possible or common,
then it is assigned the positive state. For example, tensile
force is preferred in mechanical design, because buckling
cannot occur. Conversely, fluids can support tension in only
a few, special circumstances. Fluids are usually in compression. Hence, compressive pressure is positive. The sign reversal between stress in solids and pressure in fluids is not
the end of the trouble. The sign conventions of solid mechanics and vector mechanics also differ. In vector mechanics, a
positive force is a force which acts in the positive direction,
relative to a coordinate reference frame. A force which is acting in the positive direction can either push or pull, putting
an element in either tension or compression. Added to this
confusion is the unfortunate linear graph sign convention
of assuming both power flows are in the positive direction
into the transducer. This is impossible because a transducers
only energetic property is to interface two subsystems. Ideal
transducers can neither store nor dissipate energy. All energy
which flows in must immediately flow out. This sign convention leads to the power flows having the same magnitude
but opposite signs.
When working with fluid system transducers, such as
a piston-cylinder, define your sign convention, such that a
compressive force transmitted to a fluid produces positive
pressure. Then, use PA + PB = 0 to derive the relationship between displacement and volume flow rate. This sign conven-

R winding

R Load

iG
g

R Load 2

tion avoids negative fluid pressure. Negative pressure signals possible fluid cavitation (vaporization), which is very
destructive to machines. The existence of bubbles caused by
vaporization is not the problem, although, in general, you
want to avoid creating foam or froth when pumping fluid.
The damage caused by cavitation is created by the high
stresses which occur when the bubble collapses. Unfortunately, the linear graph sign convention often leads to volume flow rate Q in a non-sensical direction. It is a no-win
situation. Resist the temptation to fix the signs until after
the system equation has been derived.

6.4.3.1 Reciprocating Piston Pumps


Reciprocating piston pumps are positive displacement
pumps. Their simple geometry allows the transducer equations to be determined from a drawing. The simplest are a
single piston with a check valves on the inlet and outlet to
maintain pressure during the pumping phase, and to prevent
backflow during the suction phase. Reciprocating piston
pumps are capable of enormous pressures. A fluid accumulator or capacitor is needed on the output to reduce the pressure
fluctuations from the reciprocating action.
6.4.3.2 Gear Pumps
A gear pump consists of two, counter-rotating, meshed spur
gears within a housing, Fig.6.55. One of the two gears is
driven by a shaft. The fluid between the gear teeth is moved
from the intake side to the discharge side of the pump by
the rotation of the gears. The meshing of the gear teeth prevents pressure driven flow from the high-pressure discharge
side to the low-pressure intake side fluid. Similarly, the tight
clearance between the outer surfaces of the gears teeth and
the housing also limits the pressure-driven back flow.
Since rotational mechanical power drives the pump, and
fluid power is delivered by the pump, a gear pump is a transducer which interfaces two dissimilar subsystems (Figs.6.56
and 6.57). The transducer equations for a gear pump are derived from the volume of fluid moved by the pump per rotation of the gears. The volume moved per revolution or cycle
is the pumps displacement:
displacement

volume
Vol
revolution

6 Power Transmission, Transformation, and Conversion

356

A = TP 1g

Qp
1g

-1g

Fig. 6.55 Gear pump. One of the two gears is shaft driven. The gears
counter-rotate. The fluid moved forward between the gears teeth and
the housing is forced out when the gears mesh

The flow rate of the pump Qp is the volume moved per revolution times the angular velocity of the gears. We need to
work in SI units for volume (m3) and in units of radians per
second for angular velocity, so we must divide the volume
per rotation by 2 rad:
Vol
1g
2 rad

The second transducer equation for a gear pump is derived


from the rotational mechanical power and fluid power flowing through the pump:

PA + PB = 0 TP 1g + QP p2 g = 0
1
TP 1g + Vol 1g p2 g = 0
2
1
TP + Vol p2 g = 0
2

B = QP p2g

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
trucks diesel engine through a split gear box, which can
transmit the engines power to either the pump or the drive
shaft for the wheels.

Qp

Qp =

Gear Pump
transducer
interface

1
TP = Vol p2 g
2

Again, note that both positive power flows are shown flowing into the transducer interface, even though one must flow
out, since the interface cannot store or dissipate energy.

6.4.3.3 Fans and Turbopumps


Fans and turbopumps are non-positive displacement, axial
flow pumps. Their transducer relationship cannot be determined by simple geometry. Their performance is best characterized experimentally. The torque driving a fan or pump
is usually unknown, but the shaft speed can be measured.
Typically, the pump is spun up to steady-state at a known
speed. Then, the pressure and quantity of fluid discharged
during a known time are recorded. An example of a powerful
turbopump is a fire engines pump. The current standards for
New York City require fire engine pumps to discharge 2,000

6.4.4Hydraulic Motors Fluid to Mechanical


Transducers
Hydraulic power is replacing many mechanical drive systems, due to the convenience of running a hydraulic line,
rather than constructing a mechanical transmission to move
power from point A to point B, and because the price of the
hydraulic components has decreased significantly. Hydraulic
motors and actuators have high power density, meaning
that they can deliver a lot of power using a small motor. Although high power actuators can be small, the entire hydraulic system can be large. The hydraulic fluid must be pressurized by the fluid power unit consisting of the prime mover,
either an electric motor or an internal combustion engine; a
gear pump; a fluid reservoir; and a means of cooling the hydraulic 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 translation) and type of motion output by the motor. Internally, the
pressurized working fluid can either turn a rotating rotor or
vane within a housing, or displace a piston within a cylinder. Externally, power is output as either a rotating shaft or
a translating rod. There is a remarkable variety of clever designs.

6.4.5Linear Hydraulic Motor or Hydraulic


Piston-Cylinder
Hydraulic power systems run at pressures which range from
750psi to 15,000psi, with 2,500psi as a typical pressure.
A hydraulic piston-cylinder, Fig.6.58, also called a hydraulic
ram, is a widely used linear actuator. It is capable of producing large forces, given the high working pressures. Importantly, hydraulic actuators can provide force without motion. The
hydraulic pump driving the actuator circulates the working
fluid back to the reservoir through a fluid resistor, when there
is no flow from the pump, allowing the pump to maintain the
rated pressure in the working fluid.

6.4Transducers

357

Vol
Q P = - _____ 1g
2 rad
1

TP

+x,v

Ap1g

QP

FP
g

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


and area A
+x,v

Piston with diameter D


and area A
+x,v

Qp

p1g

FP

FP

p1g

Qp

Vent to atmosphere
Vent to atmosphere
Piston or operating rod
Fig. 6.58 Cross-section of a single acting hydraulic cylinder and piston. An internal spring or an external mechanism is needed to move the
piston towards the hydraulic fluid port

A hydraulic piston-cylinder is a transducer, since it is


the interface between a fluid system and a translational mechanical system. A hydraulic ram is described as having high
power density, because of the large amount of power it can
transduce in its relatively small size, due to the high pressure
of the hydraulic fluid.
Although it is generally best to derive the first of the two
transducer or transformer equations using a geometric relationship between the two systems, the transducer equation
for a hydraulic piston-cylinder is an exception. Using the
force balance between the net pressure force acting on the
fluid side of the piston and the corresponding translational
force on the mechanical side, as the first of the two transducer 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, because they lead to fluid cavitation, the development and subsequent collapse of bubbles in the hydraulic fluid. Although
it would seem inconsequential, cavitation damages solid surfaces, including high strength steel, because of the extremely
high local stresses produced, when the bubbles collapse.
The free body diagram of the piston, Fig.6.59, shows
pressure acting normal to the surface of the piston on the

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
unchanged

fluid side, and the piston force in the positive direction for
translational velocity.
Equilibrium yields
Ap1g + FP = 0 FP = Ap1g
Use the power relationship for the transducer to derive the
second transducer relationship

PA + PB = 0 QP p1g + FP v2 g = 0
QP p1g Ap1g v2 g = 0
QP Av2 g = 0 QP = Av2 g
Although the signs of both transducer equations make sense
in this case, there is no reason that they should and they often
do not. If we reverse the hydraulic cylinder from left to right,
Fig.6.60, and we again define the relationship between piston force and pressure, such that pressure is positive, when
the fluid is in compression, we have a sign incongruity between volume flow rate into the cylinder, Qp, and the velocity of the piston, v2g.

6 Power Transmission, Transformation, and Conversion

358

A = QP p2g

Hydraulic Piston
transducer
interface

Fig. 6.61 Block diagram of positive power flows into a hydraulic


piston-cylinder

FP = Ap1g
1

+x,v

B = FP v3g

A p13

FP

A
g

Fig. 6.63 Double-acting hydraulic piston-cylinder. Area A > A 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
The linear graph sign convention, that both power flows
are positive into a transducer or transformer, Figs.6.61 and
6.62, leads to sign problems between the volume flow rate
into the cylinder, Qp=v2g, and the velocity of the piston, v3g.
We cant make the piston area, A, negative, so the sign of
either Qp or v2g must be inverted. Mechanical signs are often
a problem, since we have two independent sign conventions.
A positive force accelerates a mass in the positive direction,
or a positive force produces positive stress. Interfacing a mechanical system with a fluid system increases the sign difficulty, since positive pressure is compressive, and positive
normal stress is tensile. Ultimately, one has to return to the
signs defined on both the schematic and the linear graph to
understand what the sign of a variable means.
Hydraulic, pneumatic, and vacuum cylinders are widely
used in machine design, and there are many variations of
their design. A fundamental distinction is whether only one
side or both sides of the piston can be pressurized. If only
one side can be pressurized, the piston-cylinder is single
acting. If both sides of piston can be pressurized, then it is
double acting, Fig.6.63.

6.4.6 Rotational Hydraulic Motors


The four broad classes of rotational hydraulic motors design
types are vane, rotor, gear, and piston.
Vane- and rotor-type hydraulic motors The terminology
overlaps between the vane and rotor designs, since both have
rotors and vanes. Their rotors rotate from the high-pressure

Rotor

High and Low Pressure Port Pair


Fig. 6.64 Cross-section of a vane-type hydraulic motor. The vanes are
forced out of their pockets in the rotor by fluid pressure. Vane-type
hydraulic motors can produce torques of 2,000ft-lbs

inlet port towards the drain port by a drop in pressure, with


the pressure difference between the two creating a net force
on the rotor and a resulting torque. A vane motors vanes
are carried in pockets or slots in the rotor, and then forced
out against the smooth but non-cylindrical stator by a spring,
fluid pressure behind the vane, or a combination, assisted by
centrifugal force. A rotor-type motors vanes are fixed projections or lobes on both the rotor and the stator. There is
one fewer vane on the rotor than the stator. The rotors vanes
or lobes slide past the stators with the pinch point moving
along the stator, as the rotor rotates.
Both vane- and rotor-type motors are reversible. Some
designs are reversed by reversing the pump and drain connections 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,
and the side force on its shafts bearings eliminated, by having two pairs of high pressure and drain ports diametrically
opposite on the stator, Fig.6.64. A rotor-type motor with a
single high-pressure inlet is a unidirectional motor. A pair of
high-pressure inlet ports, located on either side of the diameter through the drain port, allows rotation in either direction.
An interesting hydraulic motor which combines aspects of
a rotational motor and a gear motor is a gerotor type motor

6.4Transducers

359

Two High Pressure Ports

Rotor

Stator

Roller Vane
Drain Port

Fig. 6.65 Cross-section of a rotor-type hydraulic motor

shown in Fig.6.65. A gerotor resembles a rotor-type pump, but


the fixed peak-like vanes of the stator are replaced by rollers.
Note that there are more lobes on the housing than there are on
the rotor. (The lobes in this design are steel rollers into the plane
of the view.) High-pressure fluid introduced from a port, where
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
type, because the lobes on the rotor function as the vane in a
conventional pump. The vanes contact rollers, rather than fixed
surfaces, to reduce wear.
Gear motorsUnidirectional gear motors are, essentially,
gear pumps driven by fluid pressure. They are capable of
speeds up to 10,000RPM, but have relatively low efficiency,
due to flow through the clearance between the gears and the
housing. Bi-directional gear motors need a pair of input ports
to create a preference pressure drop in either direction.
Piston hydraulic motors The operating pressure of piston
hydraulic motors ranges from 3,000 to 15,000psi, signifiFig. 6.66 A cross-section of the
rotor and output shaft and a bottom view of the rotor of a pistontype rotational hydraulic motor

Rolling-contact
bearings
Output shaft

cantly higher than the 2,000psi typical of internal combustion engines. There are an odd number of pistons in a rotor
that resembles the cylinder of a large caliber revolver handgun, Fig.6.66. The operating rods of the pistons have a ball
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 socketed plate is inclined, relative to the axis of rotor. The displacement of a piston in a pressurized cylinder increases, as
the rotor turns towards the top position, and then decreases,
as the rotor turns through the drained side. The tangential
component of the force, exerted against the angle plate by a
pressurized pistons operating rod end, creates a torque about
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 displacement) and cylinders on one side of the rotate are pressurized.
The cylinders on the other side are drained. The direction of
rotation is reversed by pressurizing the cylinders that were
drained, and draining those which were pressurized, except
the top cylinder that is always pressurized.

6.4.7Example 3: Linear Graph


of a Fourth-Order
Fluid-Mechanical System
The order of a system is determined by the number of independent energy storage elements in the system. The system
shown schematically in Fig.6.67 consists of a torque source
driving a gear pump. The gear pump has rotational inertia J,
and a by-pass flow path with resistance R1, which allows the
pump to run and maintain pressure, when there is no flow
out of the pump. The fluid discharged by the pump flows
through a fluid resistance R2 and an inertance I, before reaching a fluid capacitance (or fluid accumulator) C immediately
before a hydraulic piston-cylinder. The actuator moves mass
Angle plate rotates the output shaft
Piston rod ends captured
in ball and socket joints
These fluid ports pressurized
for counter-clockwise rotation

Rotor rotates from pressure side through


maximum piston displacement to drain side

Bottom View of Rotor

6 Power Transmission, Transformation, and Conversion

360
Fig. 6.67 Fourth-order fluidmechanical system

Hydraulic
Piston
Area A

C
T(t)

Fig. 6.68 Velocity and pressure


nodes of the fourth-order fluidmechanical system

Pump

Vol
Q P = ___ 1g
2
1

T(t) J

TP

2 R2 3 I

R1
g

R2

FA= Ap4g

I
3

QP R1

g
M. This system is fourth order, because there are four independent energy storages, J, I, C, and M.
The first step in drawing a linear graph is to find the nodes
of the across variables and identify them on the schematic of
the system. What are the across variables of the three subsystems? The across variable of the rotational mechanical
system is angular velocity . The across variable of the fluid
system is pressure p. The across variable of the translational
mechanical system is translational velocity v, Fig.6.68.
Draw the linear graph by first drawing and labeling the
nodes, Fig.6.69. Next, add the two transducer interfaces between their respective nodes. Then add the remaining elements between the corresponding nodes. Remember that
each branch in the linear graph represents a distinct power
flow. Never assign two properties to one branch, by using a
branch belonging to an element such as the inertia, fluid capacitance, or mass as one side of a transducer interface. The
two branches of the transducer interface represent the power
flow passing across the interface. They cannot represent any
other power flow in the system.

Hydraulic
Piston
Area A

T(t)

Fig. 6.69 Linear graph of the


fourth-order fluid-mechanicalsystem

R1

R2

Pump

QA

FA

6.4.8Rotational to Translational Mechanical


Transducers
6.4.8.1 Rack and Pinion
A rack is a linear gear. Pinion means small gear. A rack
and pinion is a transducer, since they interface rotational and
translational subsystems, Fig.6.70.
The analysis of a rack and pinion requires either knowing
or calculating the pitch radius of the pinion, so that the angular velocity of the pinion can be related to the translational
velocity of the rack
v2 g = rp 1g
or knowing either the circular pitch of the pinion, ppinion and
its number of teeth, Npinion, or the lineal pitch of the rack, prack
and the number of pinion teeth, so that translational displacement x and rotational displacement , and, hence, velocities
can be related:
x2 g =

p pinion N pinion
2

1g or x2 g =

prack N pinion
2

1g

6.4Transducers

361
Pin constrained by slot
to translate in contact with
the root of power screw.

+,

+x,v

-x,v

+x,v

1
2

Fig. 6.70 Rack and pinion showing positive directions for translation
and rotation

+x,v

2
__

+,

Fig. 6.72 Pin riding on the root of a power screw

n threads n revolutions n 2 rad

=
in.
in.
in.
+,

Fig. 6.71 Power screw and nut. The nut is constrained from rotation
and must translate along the power screw

The analysis then uses energy conservation expressed in


terms of power, since only the transducer attribute of the rack
and pinion is being considered, and it neither stores nor dissipates energy:

PA + PB = 0 TP 1g + FR v2 g = 0

Although it is possible to perform system dynamics calculations in US customary units, it is inadvisable, due to errors
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:
n 2 rad n 2 rad 1 in.
rad
=
N
in.
in. 0.0254 m
m
Since the power nut cannot rotate, it must translate to accommodate the rotation of the screw. The translation x of the nut
is related to the angular displacement as

TP 1g + FR rp 1g = 0
TP = rp FR

6.4.8.2 Power Screw


A power screw interfaces rotational power and translational
power. The nut is constrained from rotation and can only
translate, Fig.6.71. The relative motion of the power screw
and nut may be easier to visualize in Fig.6.72 showing a
pin, which represents a point on the thread of the power nut,
following just one axis of the helix of the root of the power
screw thread. The nut cannot rotate. It is constrained to translation. The screw cannot translate. It is constrained to rotation.
The transducer equation is based on the distance the
power screw advances the nut with one rotation. Hence, we
must know the pitch of the screw. Say screw has n threads
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
angular displacement are, in fact, dimensionless

1 m
x2 g =

N rad 1g
To express this in terms of the power variables, v and , we
differentiate both sides with respect to time
1 m d1g
=
N rad dt
dt

dx2 g

1 m
v2 g =

N rad 1g

The second transducer equation is derived from energy conservation expressed in terms of power, because, again, the
transducer represents only the transduction of the power
from one type of power (or energy domain) to the other:

PA + PB = 0 TS 1g + FN v2 g = 0
1 m
TS 1g + FN
=0
N rad 1g
1 m
TS = FN
N rad

6 Power Transmission, Transformation, and Conversion

362

Crank

Follower

r1

Cam

+,

Connecting rod

+x,v

r2

Fig. 6.73 Cam and roller follower

Power screws are designed to transduce rotational power


into translational power. The pitch of the power screw makes
them self-locking, meaning that they cannot be driven by
translating the nut. Consequently, power screws are nonlinear transducers, since power can flow from the rotational
power into translational power, but not in the reverse direction.

6.4.8.3 Cam and Follower


A cam and follower, Fig.6.73, is used to drive a translational mechanism repeatedly through identical displacements.
For example, cams are used to open the intake and exhaust
valves of automobile engines, synchronized with the motion
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
shaft, which, in turn, is driven by pistons during the combustion phase of the cycle.
The displacement of the follower equals the change in the
radius of the cam. The function of angular displacement
which describes radius of a cam, r ( ), should be a smoothing varying function. Otherwise, the profile will wear at the
discontinuities in the slope of the cams surface.
In automotive design, a tappet is used as the element in
contact with the cams surface. A particularly clever design,
which dates from the 1930s, is a hydraulic tappet, used to
reduce the effect of wear in the contact surfaces of the valve
train (consisting of the cam, tappet, lifting rod, rocker
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
in contact with the cam and lifter rods. The design can be
viewed as a fluid capacitor in series with a non-linear fluid
resistance created by a check valve. The time constant for
the tappet to extend and fill with oil (when lightly loaded)
is smaller than the time constant for the tappet to compress
and expel oil. Hydraulic tappets are now obsolete, having
been replaced by cam followers with rollers as the contact.
Although a roller-type follower is more complicated, its reduced friction increases gas mileage.

Slider

1
+,

+x,v

Fig. 6.74 Crank and slider. This is an in-line or centric crank and
slider, because the line of motion of the slider intersects the center of
rotation of the crank.

In the past, cams were relied on to perform some tasks,


such as automotive fuel injection, which are now accomplished using solenoids. The disadvantage of cams is that
they are hard automation, meaning that to change the motion or timing of a mechanism it must be redesigned and refabricated, rather than simply reprogrammed. The sophistication of feedback control in an automobile engine, and the
speed of modern power train control modules (microprocessors which control both the engine and automatic transmission), makes it inevitable that future internal combustion
engines will use solenoids, rather than cams to control the
valves, to yield greater fuel efficiency and lower pollution.

6.4.9Translational to Rotational Mechanical


Transducers
6.4.9.1 Crank and Slider
Depending on the design, a crank and slider mechanism can
either convert circular motion to reciprocating linear motion
or vice versa, Fig.6.74. The relationship between the angular
displacement of the crank and the translational displacement
of the slider is non-linear, and is described by trigonometry.
An example of a crank and slider mechanism being used to
convert translation to rotation is a piston engine, where the
reciprocating translational motion of the pistons within their
cylinders is converted to rotational motion by the crank shaft.

6.5 Multiport Transformers and Transducers


Thus far, we have considered transformers and transducers
with two ports through which energy passes from one subsystem to another. It is common in machine design to have
mechanical transformers and transducers with one input port
and many output ports. For example, the serpentine belt
drive in an automobile is driven by a pulley on the crank
shaft and, in turn, drives the water pump, the alternator, and,

6.5 Multiport Transformers and Transducers

363

PowerB

LA

LB

F(t)
Power A

Transformer
or
Transducer

often, a power steering pump and air conditioner compressor.


Extending the linear graph of a two-port transducer or
transformer element to a multiport element is straightforward. Adding a third port to the block model of a transducer, Fig.6.75, identifying that power flow as C, and, again,
recognizing that the transformer or transducer elements only
attribute is transformation or transduction (i.e., no energy
storage or dissipation) allow us to write energy conservation
at an instant, in terms of power, yielding

Power = Power

LC
Pivot

PowerC

Fig. 6.75 Three-port transformer or transducer block model showing


the positive power flows

+ PowerB + PowerC = 0

The power flow into the transformer or transducer at a


port must flow out of one or both of the other two. The bookkeeping is marginally more involved than that of a two-port
transducer, because the outward (or negative) flow of power
can divide between the available ports. The division of the
outward power flows results from either the summation of
the through variable (force or torque) at a node, or across
variable (velocity) around a loop, as illustrated in the following examples.

6.5.1Example 1: A Lever with Three


Attachments
The L lever shaped as shown in Fig.6.76 is pivoted at the
intersection of its two arms.
This system presents two challenges. First, how does one
define the positive direction? Second, how does one model
three elements acting on a lever?
Signs are a problem in system dynamics. The only constraints we must respect are the sign conventions of the linear graph method.
1. Sources raise the across variable, translational velocity in
this case, in the positive direction indicated by the arrowhead orienting the element.
2. The through variables of both branches of a transducer
or transformer are oriented towards ground, and power
flows in the two branches sum to zero.
3. In all other elements, the positive direction is arbitrary.

Fig. 6.76 Translational mechanical system

LA

LB

F(t)

2
1

LC
Pivot

K
g

Fig. 6.77 Schematic of translational mechanical system with velocity


nodes

4. Mechanical subsystems are restricted to a single direction, so that the power variables can be represented as
scalars, rather than as vectors.
Within these constraints, we are free to define the positive
direction as we wish. If the input is represented as a vector,
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
direction. We will establish signs graphically on the systems
schematic, as needed to formulate the system equation, and
then interpret the results of calculations, with reference to
the linear graph and the schematic.
Recall that levers are modeled as perfectly rigid and as
undergoing small displacements, where small means
small enough, that the difference between the arc, which
a node transverses, and its horizontal or vertical component can be neglected, allowing the motion to be considered
translational. We will establish the signs for the translational
motion of nodes on the lever to be positive, when the applied
force, F(t), acts in the direction shown, Fig.6.77. In other
words, counterclockwise rotation of the lever produces positive translational displacement of nodes on the lever.
Multiport Transformer ModelNow consider the lever.
The key is to view the system as both a mechanical system
and an energetic system, meaning that we will use equilib-

6 Power Transmission, Transformation, and Conversion

364

F
g

F
g

and LB = v2 g

rium, geometric compatibility, and conservation of energy to


develop a model. We will start with conservation of energy.
Recall that the energetic attribute of the transformer (and
its linear graph element) only transforms power variables.
There is no energy storage or dissipation in the transformation. Consequently, we can express conservation of energy
at any instant in terms of power flows. The lever divides the
power entering it between the other elements connected to it.
We will draw a linear graph of this system by following
our usual procedure. First, identify the nodes of discrete velocities on the schematic of the system, which we have already done. Next, start the linear graph by drawing and identifying the nodes, Fig.6.78. Then, add the interface branches
of the transformer, Fig.6.79. Remember that an interface has
two sides, i.e., two branches. Each interface (i.e., each pair
of interfaces branches) can only share one node, which is
ground, in this system. Finally, add the remaining elements
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
other element in the system existed. Starting with the interface between nodes 1 and 2, Fig.6.81, we first use geometric
compatibility to write an equation relating the velocities, v1g
and v2g. The rigid lever rotates with a single angular velocity
. The translational velocities vary with the distance of the
nodes from the center of rotation:

( LA + LB ) = v1g

Fig. 6.79 Transformer interface branches added between the interface


nodes of the L-shaped lever system

Fig. 6.80 Remaining elements added to the linear graph

F F(t) F

F
g

Fig. 6.78 Transformer nodes of the L-shaped lever system of Fig. 6.77

Fig. 6.81 Transformer interface between nodes 1 and 2

v1g

( LA + LB )

v2 g
LB

v1g =

LA + LB
v2 g
LB

Now, look at the linear graph. The branches on either side


of the interface define the power flows into and out of the
interface. Remember that power is the product of the through
and across variables of an element. Hence

P + P = 0 F v1g + F v2 g = 0
Eliminate the velocities by substitution:
F v1g + F v2 g = 0

L + LB
F A
v2 g + F v2 g = 0
LB

L + LB
F = A
F
LB
Repeat the process to determine the transform equations for
the interface between nodes 1 and 3, Fig.6.82.
Geometric compatibility ( LA + LB ) = v1g and LC = v3 g

v1g

( LA + LB )

v3 g
LC

v1g =

LA + LB
v3 g
LC

Energy conservation in terms of power

P + P = 0 F v1g + F v3 g = 0
Substitute to eliminate velocities

6.5 Multiport Transformers and Transducers

365

LA + L B
v1g= ____
v2g
LB

F
g

LA + L B
LA + L B
v1g= ____
v2g v1g= ____
v3g
LB
LC

Fig. 6.83 Linear graph of the L-shaped lever system, Fig. 6.77, with
transformer interfaces identified with one of the two transformer equations

L + LB
F v1g + F v3 g = 0 F A
v3 g + F v3g = 0
LC
L + LB
F = A
F
LC
The linear graph of the bent lever system is completed by
identifying the transformers with a transformer equation,
Figs.6.83 and 6.84.
There are two transformer equations for each transformer
interface. The transformer equations added to the list of energetic equations are
Leverage between nodes 1 and 2
v1g =

LA + LB
v2 g
LB

and

L + LB
F = A
F
LB

Leverage between nodes 1 and 3


v1g =

LA + LB
v3 g
LC

and

F K

Fig. 6.84 Equivalent linear graph of the L-shaped lever system,


Fig. 6.77. This linear graph has the same topology or connections
between elements as the linear graph in Fig.6.83

F F(t) F

Fig. 6.82 Transformer interface between nodes 1 and 3

F
F(t)

LA + L B
v1g= ____
v3g
LC

L + LB
F = A
F
LC

Gear A
N 2 Teeth

Pinion
N 1 Teeth

Gear B
N3 Teeth

Fig. 6.85 A pinion driving two gears

6.5.2 Example 2: A Pinion Driving Two Gears


The linear graph for the gear set shown in Fig.6.85, is drawn
by, as always, first identifying the nodes of distinct values
of the across variable angular velocity, and then adding the
elements, Fig.6.86. We have not identified what type of element is powering the pinion. It could be a torque or velocity
source, it could be a compliant shaft modeled as a spring,
or a fluid coupling modeled as a damper. (It could not be an
inertia. If you are unsure why not, then review rotational elements.) Let us keep the model general, and simply identify
the torque as TA.
Since we are dealing with a mechanical system, our
first preference is to use geometric compatibility to derive
a relationship between the angular velocities. The surface
speed of the pinion and both gears A and B is the same, or
they would not mesh. They would shear off teeth.
The number of teeth in each gear of a gear set is a measure
of its relative pitch circumference, and, hence, its relative
pitch radius:
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

6 Power Transmission, Transformation, and Conversion

366
Mesh between
Pinion and Gear A

TA
T

T
g

T
g

T
g

Fig. 6.87 Transformer interface between the pinion and gear A

surface speed = rp 1g = rA 2 g = rB 3 g

T
g

N
N1
1g and 3 g = 1 1g

N2
N3

2 g =

Now, examine the linear graph. The interfaces between the


pinion and gear A and the pinion and gear B required us to
divide the through variable torque supplied to node 1 as TA
into two torques, which we named T and T, Figs.6.87 and
6.88. The node equation for node 1 is not used to derive the
transformer equations. It is certainly essential to describe the
system and will appear in the equation list, but the transformer equations for each interface are independent of the
other elements in the system.
Recall that the transformer equations for gears can be derived from any two of the three equations, which describe
the interaction of the gears in terms of geometric compatibility, expressed as equal surface speeds. Energy conservation is expressed in terms of power flows, and equilibrium
of tooth forces. Tooth force equilibrium calculations require
the pressure angle and pitch radii of the pinion and gear and

Fig. 6.88 Transformer interface between the pinion and gear B

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 interface are equal

P + P = 0 T 1g + T 2 g = 0
Eliminate the velocities by substitution
N
T 1g + T 2 g = 0 T 1g + T 1 1g = 0
N2
T =

N1 1g = N 2 2 g = N 3 3 g
where the magnitude bars are needed, because speed is a scalar, but velocity is a vector. We must inspect the schematic
to determine the direction of rotation of the gears, yielding

Fig. 6.86 Linear graph of pinion driving gears A and B, Fig. 6.85
Mesh between
Pinion and Gear A

rA
3g = __

r C 1g

Mesh between
Pinion and Gear B

N1
T
N2

The process is repeated to derive the transformer equations


for the mesh between the pinion and gear B:

P + P = 0 T 1g + T 3 g = 0
N
T 1g + T 3 g = 0 T 1g + T 1 1g = 0
N3
T =

N1
T
N3

The two transformer equations for the two interfaces,


Fig.6.89, would be added to the list of energetic equations as
Mesh between pinion and gear A

N1
N
1g and T = 1 T

N
N

2
2

2g =

Mesh between pinion and gear B

N1
1g
N 3

3g =

and T =

N1
T
N3

6.5 Multiport Transformers and Transducers

N
2g = - __1 1g
N2
2

367

rA
2g = __
r 1g

N
3g = - __1 1g
N3
3

TA

Idler

Fig. 6.89 One of the two transformer equations for each interface is
added to identify the power transformations of the pinion driving two
gears, Fig.6.85

2g

TA
T

rA
3g = __

r C 1g

T
g

T
g

Fig. 6.91 Transformer linear graph elements of belt drive shown in


Fig.6.90

rA
2g = __
r 1g

Pulley B
Radius rB

Idler

Pulley A
Radius rA

3g

1g

Pulley C
Radius rC

vsurface

Fig. 6.90 A belt drive with one driving pulley, two driven pulleys, and
two idlers

6.5.3 Example 3: A Belt Driving Two Pulleys


The belt drive shown in Fig.6.90 has five pulleys, two of
which are idler pulleys that turn freely. Idler pulleys are used
in belt drives to direct the belt to a driven pulley and increase
the wrap or contact area between the belt and a pulley. Idler
pulleys also help to provide sufficient shear stress against the
face of the pulley in order to provide the required torque on
the pulleys shaft. Belt drives typically include a tensioner
which is a spring-loaded, pivoted arm that presses an idler
pulley against the belt.
The derivation of the transformer equations for a serpentine belt drive is very similar to a gear set. Both are rotational systems, where the geometric compatibility between
interfaces is surface speed. In a belt drive system, the surface
speed of the belt is constant along the belt. If the belt is modeled as inextensible, then it can bend but cannot stretch. The
belts surface speed equals the product of the magnitude angular velocity of a pulley times the pulleys radius:
surface speed vsurface = rA 4 g = rB 5 g = rC 6 g

Fig. 6.92 Transformer interface of pulleys A and B

where the magnitude must be used, because the direction of


the angular velocity depends on the direction of the belts
wrap around the pulley, i.e., clockwise or counterclockwise.
The angular velocities of the pulleys are related by the
inverse ratio of their radii, with the sign of the rotation of a
pulley determined by inspection of the schematic, and with
the direction of the rotation input pulley generally defined as
positive. In the belt drive above, pulleys A, B, and C rotate
in the same direction. Therefore,

4 g
rA

5 g
rB

6 g
rC

4 g =

r
rB
5 g and 4 g = C 6 g
rA
rA

The linear graph of the belt drive is drawn, as always, by


first identifying nodes of distinct values of the across variable angular velocity, as we have, then drawing the nodes,
adding the two branches of each transformer interface, and,
lastly, inserting the remaining branches, Fig.6.91. We will
let torque TA from an unidentified element drive the pulley A.
Having used geometric compatibility to derive one of the
two transformer equations for each interface, we now proceed to derive the second transformer equation for each, by
working with only one interface at a time. We will start with
the interface between pulleys A and B, Fig.6.92.

6 Power Transmission, Transformation, and Conversion

368

rA
3g = __

r C 1g

T
g

T
g

Fig. 6.93 Transformer interface of pulleys A and C

Conservation of energy across the interface between pulleys A and B, expressed instantaneously in terms of power
flows, is the sum of the products of the through and across
variables of the two branches of the interface

P + P = 0 T 4 g + T 5 g = 0
Eliminate the angular velocities by substitution
T 4 g + T 5 g = 0

T =

rB
5 g + T 5g = 0
rA

rB
T
rA

The procedure is repeated for the interface between pulleys


A and C, Fig.6.93.

P + P = 0 T 4 g + T 6 g = 0
Eliminate the angular velocities by substitution
T 4 g + T 6 g = 0

T =

rC
6 g + T 6 g = 0
rA

rC
T
rA

The transformer equations added to the energetic equations


are:
Interface between pulleys A and B

4 g =

rB
r
5 g and T = B T
rA
rA

Interface between pulleys A and C

4 g =

rC
r
6 g and T = C T
rA
rA

6.6Floating Sources, Transformers,


and Transducers
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,
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
in series are used to provide larger power variables than a
single element can, as well as to provide two or more taps
of different magnitudes of the power variable. An example of
transducers connected in series is a two-stage air compressor, in which the first stage draws air from the atmosphere
and supplies pressurized air to the second stage.
The only energetic elements which must be referenced to
a ground are the mechanical inertial elements, mass and
mass moment of inertia (or rotational inertia), and fluid capacitances. Inertial elements need to be referenced to a nonaccelerating velocity node. (By the way, the most recent estimate of your translational velocity, as you orbit the Milky
Ways galactic center, is 600,000mph). Fluid capacitances
must be referenced to the environment pressure, usually atmospheric pressure. All other energetic elements can act between any two nodes in a system.
Although we most often model source elements as referenced to (or attached to) a ground node, they need not be.
Sources, transformers, and transducers do not need to be referenced to ground, either. However, keep in mind that real
devices which act as sources, or perform energy transformation or transduction, have mass or mass moment of inertia.
These energetic attributes must be referenced to ground.
Modeling systems with floating sources, transformers, or
transducers is no more difficult than modeling systems in
which these elements are connected to ground. The assumed
positive directions of the through variable of both branches
of the transformer or transducer interface must be in the direction of a drop of their corresponding across variable, when
the input to the system is a positive step. If the through variable is in the direction of an increase of the across variable,
you are violating energy conservation. Sources are the only
energetic elements, in which the across variable increases in
the direction of the through variable flow.

6.6.1 Floating Mechanical Sources


Floating mechanical sources are very common, due to the
need to control vibrations. A typical motor mount is a bolted connection in which an elastomeric pad (i.e., rubbery

6.6 Floating Sources, Transformers, and Transducers

369

Concentric torsion spring and drag cup

b
K

T(t)

T(t)

J
Rigid shaft

Rigid mounting plate


Rigid mounting plate

Rigid shaft

Fig. 6.94 Schematic of an automobile engine modeled as a torque


source driving a rotational inertia and mounted on a concentric torsional spring and drag cup

polymer) carries some of the load in compression. A motor


mount is modeled as a spring and damper in parallel. Motor
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 steadystate reaction forces and torques of the motor created while
it powers the machine. The dead loads are not included in
the energetic model, because they have no effect on the dynamic 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
force or torque it carries. Consequently, dead loads must be
included in non-linear models, because they determine the
operating point at which the transient response occurs.
An example of a system which includes a floating source
is an automobile. If you have ever watched an automobile
engine as it is revved (or its angular velocity is accelerated), 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
mounts will isolate the passenger compartment from the vibration created by the engine, depending on whether the engine is mounted to the frame, comprised of the main structural elements, which runs the length of both sides in a rear
wheel drive automobile, or to a subframe in front wheel
drive automobiles.
The engine mounts are lumped into one concentric (or
coaxial) torsion spring K and rotational damper, or drag cup,
b, see Figs.6.94, 6.95 and 6.96. The schematic symbol represents a cross-section through a coaxial drag cup and torsional spring, but is easy to mistake for two drag cups and one
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
source.
Torsion spring K and damper b provide the reaction necessary for the torque source to apply torque to the mass moment of inertia J.

Fig. 6.95 Schematic of an automobile engine model with angular velocity nodes

T(t)

J
g

Fig. 6.96 Linear graph of the automobile engine model with floating
torque source

C2
p2(t)

Pump 2

p1(t) R1

Pump 1

C1

R2
R3

Hydraulic Piston
Area A

Fig. 6.97 Schematic of a third-order fluid-translational mechanical


system with two pumps. Pump two is floating

6.6.2 Floating and Multiple Fluid Sources


Any number of sources can be used in a machine, if they are
controlled properly. A fluid-translational mechanical system
with two pumps modeled as pressure sources is shown in
Fig.6.97. Energy can be stored independently in two fluid
capacitances and the mass, making it a third-order system,
Figs.6.98 and 6.99.

6.6.3 Floating and Multiple Electrical Sources


6.6.3.1 RC Circuit with Two Voltage Sources
Electric batteries are connected in series to increase voltage
and in parallel to increase current. Standard A, AA, AAA, C,

6 Power Transmission, Transformation, and Conversion

370

p2(t)

p1(t) 1 R1

Pump 1

Hydraulic Piston
Area A

1
+
1.5 VDC

Fig. 6.98 Schematic of a third-order fluid-translational mechanical


system with two pumps showing nodes of distinct values of pressure
and translational velocity

R2

p2(t)
1

p1(t)

R1
Switch 1

i R1

R3 C2 QP

FP

C1

1
+
1.5 VDC

Open (non-conducting)
Closed (conducting)
Closed (conducting)

Switch 1

i R1

iB

and D cell chemical batteries provide a voltage increase


from their negative terminals to positive terminals of 1.5
VDC. The schematic symbol for a battery is intended to resemble 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
series. Node 1 and node 2 can be switched independently.
Consequently, there are three possible energized configurations, Table6.1.
Each of the energized configurations yields a different
system equation. Energized configurations one and two are
simplified, by identifying the circuit elements with no current, during the transient portion of the step response, yielding 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
previously.
How do we determine its system equation for the voltage
across the capacitor, v3g? Can we use superposition? Specifi-

Switch 2

R1

g
Fig. 6.99 Linear graph of the third-order fluid-translational mechanical with two pumps

iC

Fig. 6.100 Electric circuit with two voltage sources (batteries) in series

R1
2

Table 6.1 Circuit configurations


Switch 1
Energized
configuration
One
Closed (conducting)
Two
Open (non-conducting)
Three
Closed (conducting)

FP = Ap4g
4

i R2

iB

Switch 2

iB 2

1.5 VDC

R3

C1

3 R2

Pump 2

R2

C2

iC

g
Fig. 6.101 Energized configuration one

R2

2
+

Switch 2

i R2

iB 2

1.5 VDC

1
+
1.5 VDC

3
iC

iB

g
Fig. 6.102 Energized configuration two

cally, is the step response of configuration three the sum of


the step responses of the configurations one and two? No,
summing the step responses of configurations one and two
do not yield the step response of configuration three.

6.6 Floating Sources, Transformers, and Transducers

R2

2
+

371

Switch 2

v21 + v1g v3 g
R2

i R2

1
+

R1
Switch 1

1.5 VDC

i R1

iB

iC

Fig. 6.103 Energized configuration three

To understand the response of configuration three, we


best proceed methodically and write the energetic equations.
Energetic Equations

Loop Eqs: v1g = v13 + v3 g

iR1 + iR2 = iC
v21 + v1g = v23 + v3 g

Elements Eqs: v13 = iR1 R1 v23 = iR2 R2 iC = C


Energy Eqs: E System

1
= EC = Cv32g
2

dv3 g
dt

Superimposing (or summing) the step responses of configurations one and two yields a steady-state voltage across
the capacitor of v3 g = 4.5 VDC. This cannot be the equilibrium state. If the steady-state voltage in the capacitor were
v3 g = 4.5 VDC, then there would be current flowing from the
capacitor back through both resistances in steady-state
v1g = v13 + v3 g
v1g v3 g = iR1 R1
v1g v3 g
R1

R1

iR1 + iR2 = iC

3.0 VDC 1.5 VDC


+
= iC
R1
R2

Although steady current flow through the resistances is acceptable in steady-state, the steady-state capacitor current
must be zero, or the energy in the capacitor would change,
Fig. 6.104. For the voltage across the capacitor, v3, and,
1
hence, the energy in the capacitor EC = Cv32g to be con2
stant, the steady-state capacitor current iC must be zero.
Superposition works, when additional inputs are applied
to the same input node, or driving point. Superposition
fails in this system, because the two voltage sources act on
different input nodes.
We will now reduce the equation list to derive the system
equation for the voltage across the capacitor in configuration three. The only difference between this reduction and
our previous is that we now keep two inputs. We will write
the inputs v1g and v21 as v1g (t ) and v21 (t ) to identify as inputs.
Reduction: Inputs v1g (t ) and v21 (t ) Output v3g
v21 (t ) + v1g (t ) = v23 + v3 g v21 (t ) + v1g (t ) = iR2 R2 + v3 g

v21 (t ) + v1g (t ) = iC iR1 R2 + v3 g


v21 (t ) + v1g (t ) = R2 C

v1g v3 g = v13
v1g v3 g

1.5 VDC
R2

Hence

Node Eqs: iB1 = iB2 + iR1

1.5 VDC + 1.5 VDC 4.5 VDC


= iR2
R2
iR2 =

iB 2

1.5 VDC

v21 (t ) + v1g (t ) = R2 C

= iR1

1.5 VDC 4.5 VDC


3.0 VDC
=
= iR1 iR1 =
R1
R1

v21 (t ) + v1g (t ) = R2 C

dt

dv3 g

dv3 g
dt

dv3 g

dt

iR1 R2 + v3 g

v
13 R2 + v3 g
R1

v1g (t ) v3 g

) RR

+ v3 g

dv3 g R2
R
v21 (t ) + v1g (t ) + 2 v1g (t ) = R2 C
+
v3 g + v3 g
dt R1
R1

and
v21 + v1g = v23 + v3 g v21 + v1g v3 g = v23
v21 + v1g v3 g = iR2 R2

v21 + v1g v3 g
R2

= iR2

dv3 g R2
R
v21 (t ) + 1 + 2 v1g (t ) = R2 C
+ 1 + v3 g
R1
dt
R1

dv3 g R1 + R2
R + R2
v21 (t ) + 1
v1g (t ) = R2 C
v3 g
+

dt R1
R1

6 Power Transmission, Transformation, and Conversion

372

R2

2
+

Switch 2

i R2

iB 2

1.5 VDC

1
+

Switch 1

1.5 VDC

iB

i R1

iC(ss)=0

Switch 2

i R2

iB 2

1.5 VDC

R1

R2

R1
Switch 1

i R1

Fig. 6.105 Steady-state current loop of the system in configuration


three

g
Fig. 6.104 No current flows through the capacitance during steadystate of the step response when the system is in configuration three

R1
R1 R1 + R2
R + R v21 (t ) + R + R R v1g (t )
1
2
1
2
1
R R dv3 g R1 R1 + R2
v3g
= 1 2 C
+
dt
R1 + R2
R1 + R2 R1
The System Equation is:

iR1 ( ss ) + iR 2 ( ss ) = iC ( ss )

v21 ( ss ) = v23 ( ss ) v13 ( ss )


v21 ( ss ) = iR2 ( ss ) R 2 iR 1 ( ss ) R1

RR

We have established that in steady-state iR1 ( ss ) = iR2 ( ss ).


Hence
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 )
v21 ( ss )
= iR1 ( ss )
R2 + R1

The resistance in the time constant has the same form as the
equivalent resistance of two resistors in parallel
Requiv =

R1 R 2
R1 + R 2

1
1
1
+
R1 R 2

The circumstances are similar but not identical to the equivalent resistance, which replaces two resistors connected
between the same two nodes. In the present case, the two
resistors have only node 3, where they connect to the capacitor, in common. Nevertheless, we can conclude that the time
constant of configuration three is less than the time constants
of either configurations one or two, because the resistance to
two parallel resistances is less than that of either individual
resistance.
The steady-state voltage across the capacitor can be understood, if one recognizes that absence of current into the
capacitor at node 3 yields

iR1 ( ss ) = iR 2 ( ss )

A current loop through the two resistors and battery B2 is created in steady-state, Fig.6.104.
The applied voltage drops over each resistance, in proportion to its resistance divided by the total, because the resistors have the same current. We will calculate voltage v31 ( ss )
across resistor R1 from node 3 to node 1, and add it to voltage
v1g ( ss ) across battery B1 to determine the voltage across capacitor v3 g ( ss ) in steady-state

R1
R1 R2 dv3 g
R + R v21 (t ) + v1g (t ) = R + R C dt + v3 g
1
2
1
2


= 1 2 C
R1 + R2

Therefore,
v13 = iR1 R1
v31 =

v13 = v31 = iR1 R1


R1

R 2 + R1

v21 ( ss )

When we add this result to the voltage of battery B1, v1g, it


agrees with steady-state value of the step response we derived from the system equation
R1
R + R v21 ( ss ) + v1g ( ss ) = v3 g ( ss )
1
2

6.6 Floating Sources, Transformers, and Transducers

373

3.0

LA

2.5

F(t)

Configuration 2

2.0

v3g (t)
1.5
VDC

LB

Configuration 3

1.0

Configuration 1

0.5
0.0

10

15

t, sec

20

25

Pivot

30

Fig. 6.106 Step responses for the system in Configurations 1, 2, and 3

The step responses for the system in configurations one,


two, and three, using R1 = 1, R2 = 2, and C = 3, are shown in
Fig.6.106.

Fig. 6.107 Schematic of a lever with its pivot on a shock mount

LA

LB

F(t)

Pivot

6.6.4 Floating Transformers and Transducers


So far, we have modeled transformers and transducers which
were referenced to ground. In the general case, transformers
and transducers can be located anywhere within a dynamic
system. A transformer or transducer which is not referenced
directly to ground is known as floating. The transformation
or transduction relationship is not affected. The across variable differences are those across each branch in the positive.
Regardless of where in a linear a transducer or transformer
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
variable; otherwise, you both create power as you transform
or transduce it, violating energy conservation.
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
different physical configurations and a wide range of capacities and applications in machine design. In some cases, the
device is intended simply to provide the compliance necessary to accommodate misaligned parts, which occur due to
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
shock coupling (for rotation), or, if the forces or torques are
very large, a snubber.
The mechanical schematic in Fig.6.107 shows the pivot
of a lever mounted on a spring and damper in parallel. The
linear graph for this system is drawn following the usual
method. Again:
1. Find the nodes of the across variables on the schematic,
Fig.6.108.

b
g

Fig. 6.108 Schematic of a lever with pivot on a shock mount showing


nodes of distinct translational velocities

LA
v32 = - __
v
LB 12
1
3

F
F(t)

M
b

K
g

Fig. 6.109 Linear graph of a lever with pivot on a shock mount

2. Draw and label the nodes.


3. Add both branches (sides) of transformer or transducer
elements.
4. Add the remaining elements, Fig. 6.109.
In this case, because the transformer is floating, we must
check after we add all of the remaining elements that the assumed directions for the through variables in the transformer
branches is in the direction of a drop in the across variable.

6 Power Transmission, Transformation, and Conversion

374

+x,v

F(t)

L2
V2g = _____ V1g
L 1+ L 2

b1

L1

K
M

2
L2

F(t) b1 F

b2
g

Fig. 6.110 Schematic of a translational mechanical system with a lever


showing nodes of distinct value of the across variable, translational velocity

Fig. 6.111 Linear graph of the mechanical system shown in Fig.6.110


+x,v

F(t)

6.7Equivalent Elements in Systems with


Transformers and Transducers

L1

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
dissipation element with a single equivalent element. We
must size the equivalent element, so that it will store (or dissipate) the same amount of energy (or power) as the mechanical transformer plus the element it replaces. To determine
the equivalent element, we will use an auxiliary, simplified system, consisting of just the source driving the original
system, the transformer or transducer, and the energetic element. This auxiliary system will be reduced to an equivalent
system, consisting of the source and an equivalent energetic
element. The method is best presented using an example.

L2

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.
We wish to eliminate the lever but maintain the dynamic
relationship between applied force F(t) and the velocity of
node 1, the point of application of force F(t). Consequently,
we need to determine the equivalent mass, spring constant,
and damper coefficient b2, so that when the lever is eliminated, and the elements are connected between node 1 and
ground, they provide energetic properties equivalent to the
original system. We will use an auxiliary system consisting of the force source, the lever, and one of the dynamic
elements connected between node 2 and ground in order to
determine the equivalent element parameter. We will first determine the equivalent element for damper b2.

6.7.1Equivalent Elements in a System with a


Transformer

b2 M

b2
g

Fig. 6.112 Auxiliary system constructed by removing elements from


the original system of Fig.6.110
+x,v

F(t)

b2 equiv

Fig. 6.113 Equivalent system created by removing lever and changing


the parameter value of damping b2

The auxiliary system is created by removing all of the


energetic elements from the original system, except the force
source, the lever, and damper b2, Fig.6.112. The desired
equivalent damper is shown in an equivalent system consisting of just the force source acting directly on the equivalent
damper b2equiv , Fig.6.113.
For damper b2equiv to be equivalent to damper b2 in the original system, the effect of the leverage acting on damper b2
must be included in the parameter value b2equiv. This is done by
drawing the linear graph for the auxiliary system, shown in
Fig.6.114a. Extract the energetic equations, and reduce them
to equate the force F acting through the transformer branch
on the node on one side of the interface, to a force written

6.7 Equivalent Elements in Systems with Transformers and Transducers

L2
V2g = _____ V1g
L 1+ L 2

F(t) F

b2 equiv

F(t)

in terms of the original damping coefficient b2, and the lever


ratio. This effectively removes node 2 from the system, yielding the system shown in Fig.6.114b.
If the auxiliary and equivalent systems have the same
dynamic properties, then the power flows from the forces
sources in the two systems must be identical. Consequently,
the power flows from node 1 to ground in both systems are
equal:

F(t) F

This is consistent with the objective of the equivalent system to have the same feel as the original system, since it
requires
F = Fb2

equiv

We will write the energetic equations for the auxiliary system (with a single energy storage or dissipation element).
Then, reduce either the element equation or energy storage
equation, so that it is written in terms of the power variables,
F and v1g. The term in that equation comprises of the element parameter and the transformer ratio, n, is the equivalent
parameter.
1
F = F .
n

F
g

F = n 2 b2 v1g
The equivalent damping coefficient is


b2equiv

L2
= n b2 =
b2
L1 + L2
2

Transformer Eqs: v2 g =

1
L2
v1g nv1g F = F
n
L1 + L2

Node Eqs: F (t ) = F F FK = 0
Loop Eqs: (both trivial) v1g = v1g v2 g = v2 g

Loop Eqs: (both trivial) v1g = v1g v2 g = v2 g

Element Eqs:

Reduction Start with the element equation for the damper,


since there is no energy equation. Substitute to eliminate Fb2
and v2 g
Fb2 = b2 v2 g

F = b2 nv1g

1
F = b2 nv1g
n

(6.20)

Check the physical reasonableness of this result. A larger


damping coefficient yields a larger force acting through a
damper, which is needed for a unit velocity drop across a
damper. Our equivalent damper will be connected between
node 1 and ground, and will have a larger velocity drop than
the original damper, which was connected between node 2
and ground. In order to have the same force from the source,
the equivalent damping coefficient must be smaller than the
original damping coefficient, which is our result.
The analysis is repeated for the spring and the mass,
Fig.6.115.

Node Eqs: F (t ) = F F Fb2 = 0

Element Eqs: Fb2 = b2 v2 g

We have eliminated the unwanted power variables. Now, rearrange the equation into the form of the elemental equation
of a damper

equiv

v1g nv1g
L1 + L 2

Fig. 6.115a Auxiliary system to determine the equivalent spring.


bAuxiliary system to determine the equivalent mass

F v1g = Fb2 v1g

L2

F
g

Fig. 6.114a Linear graph of the auxiliary system consisting of the


force source, the lever, and damper b2. b Linear graph of the equivalent
force source acting directly on the equivalent element b2

L2
V2g = _____ V1g
L 1+ L 2

F(t) F

Transformer Eqs: v2 g =

L2
V2g = _____ V1g
L 1+ L 2

b2

375

dv2 g
dFK
= Kv2 g FM = M
dt
dt
FK2
Energy Eqs: E Auxillary = E K E K =
2K
system
1
and E Auxillary = E M E M = Mv22g
2
System
The reduction for the equivalent spring will be demonstrated
twice: first, starting with the energy equation and next, starting with the element equation.

6 Power Transmission, Transformation, and Conversion

376

Reduction starting with the energy equation for a spring


F2
EK = K
2K

1
F
n
EK =
2K

EK

( F )
=

2K

F(t) b1

EK =

F2

2 n2 K

Mequiv Kequiv

Fig. 6.116 Equivalent system

K equiv

L2
=n K =
K
L1 + L2
2

1
d F
d F
n
dFK
= Kv2 g
= K nv1g
= nKv1g
dt
dt
dt

dF
= n 2 K v1g
dt

F(t) b1

Reduction starting with the element equation for a spring

b2

b total

F(t)

equiv

Fig. 6.117a Auxiliary system of the force source and dampers b1 and
b2 equiv. b Equivalent system of force source and damper btotal

Auxiliary system:

also yields

Node Eq: F (t ) = Fb1 + Fb2

K equiv

L2
=n K =
K
L1 + L2
2

(6.21)

The physical reasonableness check of this result parallels


the damper. Moving the spring up on the lever reduces the
mechanical advantage of the lever on the spring. The spring
must be softer, if it is higher on the lever for it to feel the
same as the force source.
The reduction for the equivalent mass will start with the
energy equation for the mass

EM =

1
1
Mv22g E M = M nv1g
2
2

equiv

Loop Eq: (trivial) v1g = v1g


Element Eqs: Fb1 = b1v1g Fb2

equiv

Equivalent system:
Node Eq: F (t ) = Ftotal
Loop Eq: (trivial) v1g = v1g

Reduction
F (t ) = Fb1 + Fb2

L2
M equiv = n 2 M =
M
L1 + L2

= n 2 b2 v1g

Element Eq: Ftotal = btotal v1g

yielding


equiv

yielding

b2

equiv

(6.22)

The last step is to draw the linear graph of the system with
equivalent elements on node 1, Fig.6.116, then combine any
like elements. One further simplification is possible. The two
dampers in parallel can be combined, Fig.6.117.

F (t ) = b1v1g + n 2 b2 v1g

F (t ) = b1 + n 2 b2 v1g
and

Ftotal = btotal v1g F (t ) = btotal v1g

Therefore,

btotal = b1 + n 2 b2

6.8 Example Problems

377

F(t)

b1+b2

equiv

Mequiv Kequiv

g
Fig. 6.118 Simplified equivalent spring, mass, and damper system

Summarizing the equivalent elements, Fig.6.118,


2

K equiv

L2
=n K =
K
L1 + L2
2

L2
M equiv = n 2 M =
M
L1 + L2
2

btotal

L2
= b1 + n b2 = b1 +
b2
L1 + L2
2

6.8 Example Problems


6.8.1Example Problem 1: Linear Graph
of a Hybrid Rotational Translational
System
A hybrid rotational and translational mechanical system is
shown schematically in Fig.6.119. The input is an angular

velocity source, (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
film with damping b2. Gear N3 engaged in the rack drives
the input shaft of a fluid coupling (drag cup) with damping
b1. The output shaft of the fluid coupling drives flywheel J.
Example Problem One Draw a linear graph of this system
and determine the transformer or transducer equations.
The key step is to identify the nodes of the across variables in the system. This system consists of three subsystems: rotational one, translational, and rotational two. The
across variables are rotational and translation velocities,
Fig.6.120.
There are two complementary perspectives on the power
flow through this multiport transducer. The first is to see
the rack as both transducing and dividing the power flowing
into the rack from the pinion. This perspective is presented
in the linear graph shown in Fig.6.121.
Note that the rack acts as a transducer in the interface between nodes 2 and 3, and as an idler in a transformer in the
interface between nodes 2 and 4.
In the second perspective, the rack is seen as transducing
power across the interface between nodes 2 and 3, distributing some of the power to mass M and damping b2, and transducing the remaining power across the interface between
nodes 3 and 4, as shown in the linear graph of Fig.6.122.
Both linear graphs are correct.
Derive the transducer and/or transformer equations. The angular velocities of the pinion and the gear are proportional to
the translational velocity of the rack. The pinion and gear are
described by their number of teeth. The rack is described by
its lineal pitch, or the number of teeth per inch, which must be
converted to teeth per meter:

Fig. 6.119 A hybrid rotational


and translational mechanical system

Compliant Shaft
Torsional Spring K

Angular Velocity Source, (t)

Pinion, N2 teeth
Rack, N1 teeth per inch
Flywheel
Rotational Inertia J

Gear, N3 teeth
Mass M
Lubricant Film
Damping b2
Fluid Coupling
Damping b1

6 Power Transmission, Transformation, and Conversion

378
Fig. 6.120 Nodes of distinct
values of the across variables,
rotational and translational velocities, shown on the schematic
of the hybrid mechanical system

Compliant Shaft
Torsional Spring K
Pinion, N2 teeth
Rack, N1 teeth per inch
Flywheel
Rotational Inertia J

2
Gear, N3 teeth

Mass M
Lubricant Film
Damping b2

Fluid Coupling
Damping b1

2 v3g= X 2g 3

TR

(t)

g
Fig. 6.121 Linear graph showing the rack dividing power between the subsystems with mass
M andmass moment of inertia J

Angular Velocity Source, (t)

4g= Y2g 4

FR M

b2 T

b1

g
Fig. 6.122 Linear graph showing the power being transferred
to the rack, and the remainder
flowing into the subsystem with
the mass moment of inertia J

(t)

2 v3g= X 2g 3

TR

FR M

4g = Z v2g 4

b2 F

b1

inch
teeth
teeth
N1
= 39.37 N1
inch 0.0254 meter
meter
The transducer constant relating the angular velocity of the
pinion in radians per second to the translational velocity of
the rack in meters per second is based on equal speeds of
their pitch surfaces:
N 2 teeth
teeth

2 rad 2 g = 39.37 N1 meter v3 g

N 2 teeth
meter
v3 g =

2 g
39.37 N1 teeth 2 rad
v3 g =

N 2 meter
2 g
N1 (39.37 )( 2 rad )

v3 g = X 2 g where

X =

N 2 meter
N1 (39.37 )( 2 rad )

6.8 Example Problems

379

Calculate the transducer equation, which relates the torque


of the pinion to the force acting on the rack:

Input Force, F(t)


L1

PA + PB = 0 TR 2 g + FR v3 g = 0

Damper b

Spring K 2
Fig. 6.123 Translational mechanical system. Force source acting on a
pivoted beam

N 3 teeth
teeth

2 rad 4 g = 39.37 N1 meter v3 g


39.37 N1 teeth 2 rad

v3g

meter
N 3 teeth

4 g =

N (39.37 )( 2 rad )
where Z = 1
N 3 meter

The corresponding transducer equation relating the force of


the rack to the torque on the gear is

Table 6.2 Parameter values


L1
L2
L3

5m

400

2m

3m

K1
N sec
m

10

K2
N
m

PE + PD = 0 T 2 g + T 4 g = 0

F v3 g + T Z v3g = 0 F = Z T

T 2 g + T Y 2 g = 0 T = YT

The transformer equation, which relates the pinions angular


velocity directly to the gears velocity, results from eliminating the racks velocity from the proceeding transducer equations:
v3 g = X 2 g and 4 g = Z v3 g 4 g = ZX 2 g
N1 (39.37 ) ( 2 rad )
N 3 meter

4 g =

N 2 meter

N1 (39.37 ) ( 2 rad )
N2
2 g
N3

4 g = Y 2 g where Y =

N2
N3

2 g

20

N
m

The transformer equation relating the pinion and gear torques


is

F v3 g + T 4 g = 0

PC + PD = 0

4 g =

pivot

L3

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:

4 g = Z v3 g

L2

Spring K1

TR 2 g + FR X 2 g = 0 TR = X FR

6.8.2Example Problem 2: A Pivoted Beam


(Lever) Acting on Three Elements
A translational system is shown in Fig.6.123. The system
consists of a rigid, massless beam supported by a pivot and
acted on by input force F(t), springs K1 and K2, and damper
b. The parameter values and the spacing of elements along
the beam are given in Table6.2. The system is in steady-state,
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
is applied to the system.
Example Problem Two, Part A Derive the system equation
that relates input force F(t) to the velocity of the location,
where spring K2 is attached to the beam.
Identify the nodes of distinct values of the across variable,
translational velocity, on Fig.6.125.

6 Power Transmission, Transformation, and Conversion

380

Po
w
er

200

F(t)
150
newtons

F(t)

100

Psource

Pspring K1

50

Pdamper

Spring K1

t, sec

Pspring K2

Damper b

Fig. 6.124 Force applied to the system shown in Fig.6.123


Spring K2

Input Force, F(t)

Fig. 6.126 Power from force source viewed as flowing into the lever,
and then dividing between the two springs and the damper

L1

1
Spring K1

L2

v 2g =

pivot

L2
____

v1g

L1 +L 2

v 3g =

-L 3
____

L1 +L 2

v1g

L3

Damper b

Spring K 2

Fig. 6.125 Nodes of distinct values of the across variable, translational


velocity

Three equivalent linear graphs are shown. The first two


have identical topology, or connectedness, and represent
power from the source dividing between spring K1, at node
1, damper b, at node 2, and spring K2 at node 3, Figs.6.126,
6.127 and 6.128. The division of power is represented by the
product of the velocity of node 1 times the continuity equation for node 1:

F (t ) v1g = FK1 + F + F v1g


F (t ) v1g = FK1 v1g + F v1g + F v1g
where F v1g and F v1g are the power flows to the damper and
spring K2, respectively.
An alternative linear graph divides the power from the
source at node 1 between spring K1 and node 2. The power
transmitted to node 2 then divides between the damper and
node 2, Figs.6.129 and 6.130.
Transformer Equations: Use geometry to derive the first
transformer equation of each pair, then sum power flows
down the branches of the transformer interface to derive the
second transformer equation.

F F(t) K1 F

K2

Fig. 6.127 Linear graph showing power dividing where it enters the
beam, at node 1

F(t)

v 2g =

L2
____

L1 +L 2

K1 F

v 3g =

v1g 2

-L 3
____

L1 +L 2

b F

v1g

K2

Fig. 6.128 Linear graph with same topology of Fig.6.127

Geometry: There is one angular velocity of the rigid


beam. All points on the beam have the same angular velocity
but different translational velocities

beam =
v2 g =

v1g
L1 + L 2

L2
v1g
L1 + L2

v2 g
L2

and v3 g =

v3 g
L3
L3
v1g
L1 + L2

6.8 Example Problems

381

er
Pow

Name the lever ratios, A and B, to reduce effort and the likelihood of transcription error:

F(t)

Psource
Pspring K1

Po
we
r

Pnode 2

Spring K1

Pnode 3

Damper b

Pspring K 2

dF (t )
dt
dF (t )
dt

L3
L3
v1g = 0 F =
F
L1 + L2
L1 + L2

The derivation of the system equation for the velocity of the


beam, at the point where spring K2 is attached, will use the
linear graph of Fig.6.127.
The system has two springs but only one independent energy storage mode. Given the deformation of one spring, the
deformation of the other can be calculated, since the beam is
rigid. The system equation will be first-order.

Continuity Eqs:
v2 g = v2 g
dt

= K1v1g

FK2 F = 0
v3 g = v3 g
dFK2
dt

v2 g =

L2
v1g
L1 + L2

F =

L2
F
L1 + L2

v3 g =

L3
v1g
L1 + L2

F =

L3
F
L1 + L2

FK21
2 K1

EK =
2

FK22
2K2

F (t ) = FK1 A F + B F

= K 2 v3 g .

dv1g
K1
v3 g + A2 b
B K 2 v3 g
B
dt

1 2 dv3 g K1

+ B K 2 v3 g
A b

B
dt
B

dv3 g
dF (t )
= A2 b
+ K1 + B 2 K 2 v3 g
dt
dt

dv3 g
dF (t )
B
A2 b
=
+ v3 g
2
2
K1 + B K 2 dt
K1 + B K 2 dt

Substitute in the lever ratios

Energetic Equations

Fb F = 0

F = BF .

dv2 g
dF (t )
= K1v1g + Ab
B K 2 v3 g
dt
dt

P + P = 0 F v1g + F v3 g = 0

dFK1

v3 g = Bv1g

dFK2
dv2 g
dF (t ) dFK1
=
+ Ab
B
dt
dt
dt
dt

L2
L2
F
v1g = 0 F =
L1 + L2
L1 + L2

Element Eqs: Fb = bv2 g

L3
L1 + L2

F = AF

F (t ) = FK1 + A Fb B FK2

P + P = 0 F v1g + F v2 g = 0

Compatibility Eqs: v1g = v1g

F (t ) = FK1 + F + F

Summation of transformer power flows

F (t ) = FK1 + F + F

v2 g = Av1g

Reduction: Input F(t), Output v3g

Fig. 6.129 The alternative is to view power from the source flowing
into node 1, the remaining power to node 2, and then what remains
flowing to node 3

F v1g + F

L2
L1 + L2

Energy Eqs: E sys = E K1 + E K2 E K1 =

Spring K2

F v1g + F

L3
L1 + L2
2

L3
K1 +
K2
L1 + L2

dF (t )
=
dt

L2
b
L1 + L2
2

dv3 g

dt
L3
K1 +
K2

L1 + L2

+ v3 g

Check the units of the system equation in terms of the power


variables and time. Transformer constants are dimensionless.
The units of the element parameters are
F

[ K ] = t v [b] = v

6 Power Transmission, Transformation, and Conversion

382

v 2g =

L2
____

L1 +L 2

v1g

-L 3
____

v 3g =

L1 +L 2

v1g

Input Force, F(t)

F(t)

K1 F

F b F

K2

Spring K1

L1+L2
pivot

L3

Fig. 6.130 The linear graph corresponding to the power flows shown
in Fig.6.129

L3

L + L
dF (t )

1
2
2

dt
K + L3 K

1
2

L1 + L2
L 2

2
b

dv3 g
L1 + L2
=
+ v3 g
2
dt

L
3
K +

K
1 L1 + L2 2

1 F b v
t v F F t v v
K t = K t + [ v ] F t = v F t + [ v ]
t v F F t v v

=
+ [v]
F t v F t

[v] = [v ] + [v ]

The units check.


There are two springs in this system. Should the system
equation be second order? No. This is a first-order system,
because the deformations and forces of the two springs are
related through the rigid beam. The state variable of the system is established by finding the single equivalent spring,
which stores the same energy as the two springs in the system. We will find the single equivalent spring for node 1. The
auxiliary system is shown in Fig.6.131 and its linear graph
in Fig.6.132.
Express the energy equation for spring K2 in terms of
force F to eliminate node 3 from the auxiliary system

EK =
2

EK =
2

FK22
2K2

F
B
2K2

EK =

EK =

F
2K2

F2
2B2 K2

Spring K 2

Fig. 6.131 Auxiliary system for determining equivalent spring

v 3g =

F(t) K1 F

-L 3
____

L1 +L 2

v1g

K2

Fig. 6.132 Linear graph of the auxiliary system shown in Fig.6.131

The spring constant for the equivalent of spring K2 attached


to node 1 is
2

K equiv = B 2 K 2

L3
K2
K equiv =
L1 + L2

The last step is to remember how springs in parallel combine,


Fig.6.133. Is the equivalent spring constant stiffer or softer
than one of the original spring constants? It is stiffer, because
the two springs divide the load leading to less deformation.
Hence, the spring constants of springs in parallel sum.
2

L3
K2
K total = K1 +
L1 + L2
Example Problem Two, Part B Determine the velocity of
the location where spring K2 is attached to the beam, as a
function of time in response to the input plotted in Fig.6.124.
Use Mathcad or MATLAB to plot that velocity.
Use superposition of the unit step response to formulate
the response function. The system is assumed to be de-energized, prior to the application of the first step input at an
unknown time but long enough prior to time t=0, that the

6.8 Example Problems

383

L3

( )

v3 g 0+ =

F(t) K1

F(t)

K equiv

K total

( )=

v3 g 0

Fig. 6.133a Linear graph of the auxiliary system with the equivalent
spring in parallel with spring K1. b Equivalent spring combined with
spring K1

system has reached steady-state by time t=0. We will assume that the first step input was applied at time t= 7,
seven time constants before the pulse is applied.
Identify the type of step response from the initial and
steady-state values. To use superposition we must assume
that the system is de-energized, prior to the input of the
unit step. Determine the value of the output variable at time
t = 0+ , the instant after a unit step input was applied to the
de-energized system.
Since the system is assumed to be de-energized before
the application of the unit step input, both spring forces are
zero. The forces acting through the springs cannot change
instantaneously, because springs store energy. The applied
unit step must be picked up by the damper at time t = 0+. The
transient response of the system is due to the transfer of the
force from the damper to the two springs.
Calculate the force acting through damper b in order to
determine the velocity of node 2 at time t = 0+ , v2 g (0+ ). Find
v3 g (0+ ) from v2 g (0+ ).

( ) = 1N = F (0 )

F 0

( )
+

K1

( )

1N = F 0 + F 0

( )

+ F 0 + F 0
+

( )

( )

( )

Fb 0+ =

( )

bv2 g 0+ =

1
1N
A

( )

v3 g 0+ =

( )

1 N = AFb 0+

1
1N
A

( )

v1g 0+ =
B
1N
bA 2

1
1N
bA 2

L 3 L1 + L 2
b L 22

1N

1N

L3
K1 +
K2
L1 + L2

dus ( ss )
dt 0
2

L2
L + L b
1
2

dv3 g ( ss )

dt

L3
K1 +
K2
L1 + L2

+ v3 g ( ss )

v3 g ( ss ) = 0

L3
L + L
1
2
2

L3
K1 +
K2
L1 + L2

dF (t )
=
dt

L2
L + L b
1
2

dv3 g

dt
L3
K1 +
K2

L + L2
1

+ v3 g

( )

1N = AF 0 + BF 0

1 N = AFb 0+ B FK2 0+

L3
L + L
1
2

( )

L2
b

L1 + L 2

Determine the final value of the unit step response, and the
time constant from the system equation.

( )

L1 + L 2

L2
L + L b
1
2
2

L3
K1 +
K2
L1 + L2

Notice that the denominator of the time constant is the combined spring constant, Ktotal.
The unit step response is a decay to zero of the form

( )

v3 g (t ) = v3 g 0+ e
Evaluate the constants, v3g(0+) and ,

( )

v3 g 0+ =
u .s.

3m (5 m + 2 m )
m
1 N = 0.0919
N sec
sec
400
( 2 m )2
m
2

6 Power Transmission, Transformation, and Conversion

384

200
150

50 N u s(t+7)

F(t)
newtons

100 N u s (t-1)

(t 1)

m 2.39
+ 100 0.0919
e
us (t 1)
sec

100
50

-7

(t + (7 )( 2.39))

m
e 2.39 us (t + ( 7 )( 2.39))
v3 g (t ) = 50 0.0919
sec

-50

t, sec

(t 2 )

m 2.39
150 0.0919
e
us (t 2)
sec

-100

6.8.3Example Problem 3: A Serpentine Belt


Driving Two Elements

-150

-150 N u s(t-2)

-200

Fig. 6.134 Superposition of scaled and time shifted Heaviside step inputs to create the input applied to the system

8
4

v3g(t)
m 0
___
sec
-4
-8
-20

-15

-10

-5

t, sec

10

15

Fig. 6.135 Response function v3g(t)


2

2m
N sec
5 m + 2 m 400 m
2

N 3m
N
10 +
20

m 5m + 2 m
m

= 2.39 sec

The unit step response is


v3 g (t ) = 0.0919

t
m 2.39
e
sec

Create the input function by superimposing scaled and timeshifted Heaviside step functions, Figs.6.124 and 6.134.
F (t ) = 50 N us (t + 7 ) + 100 N us (t 1) 150 N us (t 2)
Construct the response function, by weighting (scaling) the
unit step response with the magnitude of the input. Timeshift the unit step response with the same time shift as the
input term. Multiply by the unit step with that time shift to
zero-out the term, until time has advanced to that term:

A rotational mechanical system is shown in Fig.6.136. The


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 rotational damper (drag cup), b = 2N m sec, and a flywheel,
J = 8kg m 2. A 5-in. diameter pulley drives the shaft attached
to the rotational dampers drag. The shaft on the rotational
dampers cup is rigidly attached to the machines frame. The
flywheel is driven by a pulley, 6-in. in diameter. There are
three idler pulleys to route the belt, and provide sufficient
wrap of the belt around the motor, damper, and flywheel pulleys. Assume that the belt does not slip.
Example Problem 3, Part A Derive the system equation
which relates the input torque to angular velocity of the
motor shaft. Check the units of the system equation in terms
of the power variables and time.
The first step is to find the nodes of the across variable
on either end of the energetic elements and identify them on
the schematic, Fig.6.137. The idler pulleys route the belt but
have no energetic properties and should be neglected.
There are two equivalent linear graphs for this system
which differ on how one perceives the power flow. First, the
idler pulleys have no effect on the power flow in the system.
They can be ignored. Second, the belt has no mechanical
properties, other than being described as not slipping on the
pulleys. No mechanical properties means the belt does not
store or dissipate energy. It simply transmits power. An ideal
belt is perfectly flexible but inextensible. The following are
the two equivalent models.
Model One. Ignore the ideal belt and view power as dividing at the torque source into two power flows. One power
flow occurs between the motor shaft and the damper shaft.
The other power flow occurs between the motor shaft and
the flywheel shaft, Fig.6.138.
Model Two. Include the ideal belt, and view the power
from the torque source as flowing into the belt, and then
dividing into two power flows at the damper shaft, one
which flows into the damper and the other into the flywheel,
Fig.6.139.

6.8 Example Problems

385

Fig. 6.136 Rotational system


driven by a serpentine belt

Flywheel
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


values of the across variable
angular velocity

Flywheel
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. 6.138a Model One. Power


from the source divides the
motor shaft. b Linear graph of
Model One

Pflywheel

Pdamper
Psource

D
D
__
2g = __4 1g 3g = 4 1g
D6
D5
1
2
3

T
g

T T(t) T
g

T
g

6 Power Transmission, Transformation, and Conversion

386
Fig. 6.139a Model Two. Power
delivered to the belt and divided
at the damper shaft. b Linear
graph of Model Two

D
2g = __4 1g
D5

Pflywheel

T(t) T

Pdamper

D
3g = __5 2g
D6

b T

Pbelt = Psource
We will work the problem using the linear graph for
Model One, Fig.6.138.
Transformer EquationsThe transformer equations are
derived by using the geometry of the system to establish
one of the two transformer equations for each interface. The
second transformer equation for each interface is established
by summing the power flows down the two branches of the
interface and equating the sum to zero. The geometric property of the system used to establish the transformer ratios is
the belt speed, vbelt. An ideal belt does not slip on a pulley.
Therefore, the surface of the pulley and the belt have the
same speed. The angular velocity of the shaft is belt speed
divided by the radius of the pulley. Hence, the geometry
yields the angular velocity relationship. Then the angular
velocity relationship and the power summation yield the
relationship between the torques.
Geometry:

D
D
D4
1g = vbelt = 5 2 g 1g = 5 2 g
2
2
D4

D
D
D4
1g = vbelt = 6 3 g 1g = 6 3 g.
2
2
D4
Power: T 1g + T 2 g = 0 and T 1g + T 3 g = 0
Torques:
T 1g + T 2 g = 0

T =

D6
1g + T 3 g = 0
D4

T =

D5
2 g + T 2 g = 0
D4

D5
T
D4
T

D6
3 g + T 3g = 0
D4

D6
T
D4

Energetic Equations
Nodes: T = T + T

T Tb = 0 T TJ = 0

Loops: 1g = 1g

2g = 2g

Elements: Tb = b 2 g

TJ = J

1 g =

D5
D
2 g T = 5 T
D4
D4

3g = 3g

d 3g
dt

1 g =

D6
D
3 g T = 6 T
D4
D4

Define temporary variable names for the pulley diameter


ratios:
A

D5
D4

and B

1g = A 2 g T = AT
Energy: E sys = E J

EJ =

D6
D4

1g = B 3 g T = BT

1
J 32g
2

Reduction: Input T, Output 1g


T = T + T
T=

T=

1
1
T T
A
B

T=

1
1
Tb + TJ
A
B

1
1 d 3g
1 1
1 1 d 1 g
b 2 g + J
T=
b1g +
J
A
B
dt
AA
BB
dt
2
2
d 1g
1
1
T = b 1g + J
A
B
dt

T=

1 d 1g
1
J
+ 2 b 1g
dt
B2
A

A2
A2 J d 1g
T= 2
+ 1g System equation
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,


1N-m. We will then construct the input function by scaling
and timeshifting Heaviside step functions. Finally, we will
construct the response function by scaling and time shifting
the unit step response with the same scaling and time shifting
used to construct the input function. We must also multiply
each term in the response function by a Heaviside unit step
function with the time shift of the input, so that each term
will be zeroed-out until its corresponding input step turns on.

T(t)
N m 150
100
50
10

-50

20

t, sec

30

-100
-150
Fig. 6.140 Input torque
2

2
D D4 J d 1g
1 D5
+ 1g
T = 5

b D4
D4 D6 b dt

D J d 1 g
1 D5
+ 1g System equation
T = 5
b D4
D6 b dt

D5 J
D b . Time constant
6

Unit Step Response Determine the value of the output variable 1g, at time t = 0+ . The energy stored in an element cannot change instantaneously without an infinite power flow.
Hence, the state variable must remain unchanged, when a
step input is applied to the system. The state variable of this
first-order system is the angular velocity of the flywheel, 3g.

E sys ( 0 ) = 0 = E sys ( 0+ )

E sys =

1
J 23 g
2

3 g ( 0 ) = 0 = 3 g (0+ )
We have assumed that the belt does not slip. Consequently,
the angular velocity of all the pulleys is proportional, including the pulley on the torque source. If the flywheels pulley is
not rotating at time t = 0+ , then neither is the torque source:

3 g (0+ ) = 0 1g (0+ ) = 0

Check the consistency of the units.

We now determine the steady-state value of output variable

D 2 1 D 2 J d 1g
+ 1g
5 T = 5
D4 b D6 b dt

1g in response to a unit step input from the system equation:

T J
b = b t + [ ]

D J d 1g ( ss )
1 D5
1 N m us ( ss ) = 5
+ 1g ( ss )

1
b D4
dt
D6 b
0

T t
+ [ ]
T
=
T T t

[ ] = [ ] + [ ]

The units are consistent.


Example Problem 3, Part BThe system is running in
steady-state under the action of a step input of torque,
T = 100 N m, applied at an unknown previous time
when, at time t = 0, the applied torque is increased to
T = +50 N m . Ten seconds later, the torque is increased
again to T = +200 N m. The input torque history is shown
in the plot of Fig.6.140. Determine the angular velocity
of the motors shaft as a function of time. Use Mathcad or
MATLAB to plot the angular velocity of the motors shaft
from 0 to t = 10 + 6 sec.
We will determine the response of the system to this input
using superposition. First, we will establish the response

1Nm

rad

1g ( ss ) =
1g ( ss ) = 0.78
4 2 N m sec
sec
u .s.
u .s.
2

D J 5 8
=
= 2.78 sec
= 5
D6 b 6 2
The first-order step response is a stable growth from an initial
value, 1g (0+ ) = 0, of the form

1g (t ) = 1g ( ss ) 1 e

1g (t ) = 0.78
u .s.

rad
2.78
1
e

sec

6 Power Transmission, Transformation, and Conversion

388

150 N m us (t)

200

T(t)
N m 150

Fluid Coupling
Damping b

100

150 N m us (t-10)

50
-7

Ideal Frictionless
Bearing

10

-50

30

20

t, sec

-100

-100 N us (t+7)

-150

N1

Angular
Velocity
Input (t)

N3

N2
N4

Inertia J

Gear 2 and Gear 3 are keyed


(fixed) to the same shaft.

Fig. 6.143 Rotational system with compound (cluster) gears. The


input is angular velocity

Fig. 6.141 Torque input

1g (t ) = 78

200
150

+ 117

1g (t) 100
rad
___
sec

50

(t +19.5)
2.78

us (t + 19.5)

rad
2.78

1
e

us (t )
sec

+ 117

0
-50
-100

rad
1 e
sec

rad
1 e
sec

(t 10)
2.78

us (t 10)

6.8.4Example Problem 4: Rotational System


with Compound Gears
-20

-10

Fig. 6.142 Response function


shaft

20

10

t, sec

30

1g , the angular velocity of the motors

Construct the input function from scaled and time shifted


Heaviside unit step functions (Figs.6.141 and 6.142):
T (t ) = 100 N m us (t + 7 ) + 150 N m us (t )
+ 150 N m us (t 10)

Construct the response function with the same scaling and


time shifting as the input function. Multiply each term of the
response function by the corresponding unit step function
from the input function

1g (t ) = 100 0.78

(t + 7 )

rad

1
e
us (t + 7 )

sec

rad

+ 150 0.78
1 e us (t )

sec

rad

+ 150 0.78
1 e

sec

(t 10)

us (t 10)

A rotational mechanical system is shown in Fig.6.143.


The angular velocity source acts on the input shaft of
a fluid coupling, modeled as a drag cup with damping
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
shaft as gear three. Gear three with N 3 = 48 teeth drives gear
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
an ideal frictionless bearing.
Example Problem Four, Part AUse the linear graph
method to derive the system equation, which relates the
angular velocity input to the torque through the fluid coupling, damping b.
Identify nodes of distinct values of the across variable,
angular velocity, on the schematic of the system, Fig.6.144.
Note that gears two and three have the same angular velocity, and the only energetic property of any of the gears is
the transformation of the power variables. Real gears have
inertia, and gear sets have significant friction. In this model,
the equivalent inertia of the gear set would be included in rotational inertia J, and the linear friction of the gear set would
be added to the damping of fluid coupling b.
The most common error is to assign two energetic properties to a branch of a transformer or transducer. Avoid this

6.8 Example Problems

389

Fluid Coupling
Damping b

N1

N3

Angular
Velocity
Input (t)

using geometric compatibility, where number of teeth on a


gear is proportional to its pitch circumference

Ideal Frictionless
Bearing

N1 2 g = vtooth = N 2 3 g

N2

N4

Inertia J

Gear 2 and Gear 3 are keyed


(fixed) to the same shaft.

Fig. 6.144 System schematic annotated to show nodes of distinct values of the across variable, angular velocity

(t)

TT

(t)

N 4 4 g = vtooth = N 3 3 g

N3
Y
N4

N3
3g
N4

4 g = Y 3 g

PA + PB = 0 T 2 g + T 3 g = 0
Substitute to eliminate 3g

T 2 g + T X 2 g = 0 T = XT
Similarly,

4g =

PC + PD = 0 T 3 g + T 4 g = 0

4g= - Z 2g

3g = X 2 g

and

N1
2g
N2

Sum the power flowing into the interface of the transformer


to establish the relationship between the torques

Fig. 6.145 Linear graph including angular velocity of Gears 2 and 3, 3g

N
Simplify notation. Define X 1
N
2
Likewise

3g= - X 2g 4g= - Y 3g

3g =

T 3 g + T Y 3 g = 0 T = YT

Energetic Equations
Continuity:

Fig. 6.146 Linear graph of equivalent system eliminating node 3

error by drawing the interface elements and identifying their


through variables, before drawing any other elements. A
branch identified by a through variable is less likely to have
a parameter assigned to it. Name the torques acting through
the gear set transformer elements as you please, but keep it
simple, Fig.6.145.
Note that there is no energetic element connected to
node3, other than the branches of two transformer interfaces. Consequently, this linear graph can be simplified, by
eliminating node 3, and expressing the velocity of node 4 as
a function of the velocity of node 2, Fig. 6.146.
We will work with the linear graph, which includes
node3. Write the transformer equations for the gear set,

TSource = Tb

Tb = T

T T = 0

T TJ = 0

Compatibility:

1g = 12 + 2 g
Elements: Tb = b 12

3g = 3g
TJ = J

3 g = X 2 g T = XT
Energy: E sys = E J

EJ =

d 4g
dt

4g = 4g
N1
X
N2

N3
Y
N4

4 g = Y 3 g T = YT

1
J 32g
2

Reduction: Input (t ) = 1g , Output Tb

= 12 + 2 g =

Tb
T
1
+ 2 g = b 3 g
b
b X

6 Power Transmission, Transformation, and Conversion

390

Tb 1

b X

4 g
Y

d 1 dTb
1 TJ
=
+
dt
b dt
XY J

d 1 dTb
1 d 4 g
=
+
dt
b dt
XY dt

Determine the unit step response. First establish the value


of the state variable at time t = 0+, assuming the system is
initially de-energized:

d 1 dTb
1
=

T
dt
b dt
XYJ

d 1 dTb
1 T
=

dt
b dt
XYJ Y

d 1 dTb
1
T
=
+
dt
b dt
XY 2 J

d 1 dTb
1 T
=
+

dt
b dt
XY 2 J X

d 1 dTb
1
=
+ 2 2 Tb
dt
b dt
X Y J

N1 N 3
d N1 N 3 J dTb
N N J dt = N N b dt + Tb
2
4
2
4
2

E J ( 0+ ) =

( 0+ ) = 1

N1 N 3
d N1 N 3 J dTb
N N J dt = N N b dt + Tb . System equation
2 4
2 4

Check the units of the system equation in terms of the power


variables and time:
T

T=J

d
dt

Tt

[ J ] =

N N 2 d N N 2 J dT
b
1 3 J
= 1 3
+ [Tb ]
N
N
dt
N
N
b
dt

2 4
2 4

J T
J t = b t + [T ]
T t T t T

=
+ [T ]
t T t

rad
= 12 0+ + 2 g 0+
sec

( )

( )

( )

+
rad Tb 0
1
=
+
4 g 0+
b
XY
sec

( )

rad
Tb 0+ = b 1
sec

( )

N m sec rad

Tb 0+ = 150
1

rad sec

( )

( )

Tb 0+ = 150 N m

Find the time constant and the steady-state value of the unit
step response from the system equation:
2

[T ] = [T ] + [T ]

( )

( )

N1 N 3
d N1 N 3 J dTb
N N J dt = N N b dt + Tb
2 4
2 4


NN J

= 1 3
N2 N4 b

The units check.


Example Problem Four, Part B The system was de-energized before the angular velocity pulse, shown in Fig.6.147,
was applied to the system. Determine the response function
for torque through the fluid coupling (damping b), for the
input plotted in Fig.6.147.
We will solve using superposition of scaled and timeshifted unit step responses. Construct the input function from
scaled and time shifted Heaviside unit steps, Fig.6.148:

(t ) = 20

( )

+
rad T b 0
1
=
3 g 0+
1
sec
b
X

[b] =

( )

4 g ( 0+ ) = 0

T = b

1
J 42g 0+ = 0 4 g 0+ = 0
2

The unit step is (0+ ) = 1(rad/sec) us (0+ ) = 1(rad/sec) . Use


the values of the input and the state variable at time t = 0+ ,
with algebraic equations of the equation list, to determine the
value of the output variable at time t = 0+ :

d
J dTb
X Y J
= X 2Y 2
+ Tb
dt
b dt
2

E sys ( 0 ) = 0 = E sys ( 0+ ) = E J ( 0+ )

rad
rad
us (t ) 20
us (t 0.5)
sec
sec

5448

1812

0.3 kg m 2
= 0.288 sec
N m sec
150
rad

N1 N 3
d rad

N N J dt 1 sec us ( ss )
2 4

N N J dTb ( ss )
= 1 3
+ Tb ( ss )
dt 0
N2 N4 b
2

Tb ( ss ) = 0

6.8 Example Problems

391

Identify the type of first-order unit step response, Tb ( ss ) = 0.


The unit step response is a decay to zero of the form

( )

Tb (t ) = Tb 0+ e

PA + PB = 0 v2 g iM + TM 3 g = 0

Tb (t ) = 150 N m e 0.288

Construct the response function, Fig.6.149, with the same


scaling and time shifting as the input function,

(t ) = 20

TM = iM

rad
rad
us (t ) 20
us (t 0.5)
sec
sec

(150 N m ) e

0.288

rad
m

and

3g

TS v4 g + FN v4 g = 0

isource = iR iR = iM

3g = 3g

Compatibility: v1g = v12 + v2 g

v4 g = v4 g

Elements:
iR = Rv12

TJ = J

TM = iM

3g =

d 3g
dt
1

v2 g

Energy: E Sys = E J + E M

dv4 g

FM = M

TS =

EJ =

Fb = bv4 g

dt
FN

1
J 32g
2

3 g = v4 g
EM =

1
Mv42g
2

Reduction: Input v1g, Output v4g


v = RiR 3 g

v = RiR v4 g

v=

204 N m
Nm
= 6.8
30A
A

FN

TM TJ TS = 0 FN FM Fb = 0

12 in 0.0254 m 1 N
150 ft lb
= 204 N m
1 ft 1 in 0.224 lb

Continuity:

TM
iM

Energetic Equations

v1g v = v12 + v2 g

TS =

Electric Motor We know the motor produces 150 ft lbs of


torque for 30 amps of current. Convert the torque to SI

Calculate transducer constant

3 g = v4 g

v4 g

PC + PD = 0 TS 3 g + FN v4 g = 0

6.8.5Example Problem 5: Hybrid Electric,


Rotational, and Translational System

Example Problem Five, Part AUse the linear graph


method to derive the system equation, which relates input
voltage to the velocity of the mass.
Draw the linear graph. The essential first step is to identify
the nodes of distinct values of the across variable in the energetic model of the system, Fig.6.151. There are three subsystems: electrical, rotational, and translational, Fig.6.152.
The energetic properties of the system are described in the
problem statement and called-out on the drawing. Do not
add additional energetic properties to the model.

v2 g

threads 1rev 2 rad 39.36 in


rad



= 495
inch 1thread
rev
1meter
m

= 495

20 us (t 0.5)

The system shown in Fig.6.150 is powered by a voltage


source which drives an electric motor. The electric motors
winding resistance is R = 8 . The mass moment of inertia
of the motor is negligible compared to the motors load. The
motor produces 150 ft lbs of torque for 30amps of current.
The shaft of the motor turns a flywheel with mass moment
of inertia J = 3 kg m 2 The flywheels shaft turns a power
screw with a linear pitch of two threads per inch. The power
screw pushes mass, M = 20 kg, on lubricated ways. The lubricating film between mass M and the ways has damping
b = 2 N sec/m .

Power Screw We know that the lineal pitch of the power


screw is 2 threads per inch. One thread equals 2 rad of rotation:

Tb (t ) = (150 N m ) e 0.288 20 us (t )
(t 0.5)

v2 g iM + iM 3 g = 0 3 g =

v=

( TJ TS ) v4 g

v=

TM v4 g
R

TJ

TS v4 g

6 Power Transmission, Transformation, and Conversion

392

v=
v=

RJ dv4 g
R
+
( FM Fb ) v4 g
dt

(t) 20
rad
___
sec

dv4 g

RJ dv4 g
R
+
M
bv4 g v4 g

dt

dt

10
0

RJ dv4 g RM dv4 g Rb
v=

v v4 g
dt
dt
4 g
RJ RM dv4 g Rb

v =
+
+
+ v4 g

dt

10

(t)
rad
___ 0
sec

v2 g

[ ] = i

[ ] =


volt

3 g = v4 g

[ ] = v

TJ = J
FM = M

d 3g
dt
dv4 g
dt

Fb = bv4 g

1,000
0
-1,000
-2,000
-3,000

0.5

1.0

Tt
[ J ] =

[ M ] =

Ft
v

[b] = v

1.5

t, sec

2.0

Fig. 6.149 Response function, the torque acting through the damper
DC Motor

volt
i

Tb(t)
N m

[ R ] =

rad
___
-20 sec us (t-0.5)

[ ] = T

v12 = RiR

t, sec

2,000

1.0

3,000

FN

TS =

0.75

Fig. 6.148 Two scaled step functions, one of which is time shifted, that
superpose to form the input pulse

Check the units of the system equation:

3 g =

0.5

-20

System equation in time constant form

TM = iM

0.25

-10

RJ + RM dv4 g

v=
+ v4 g

2 2
2 2
Rb +

dt
Rb +

RJ 2 + RM dv4 g
v=
+ v4 g
2 2
2 2
dt

Rb +
Rb
+

1.0

rad
___ u (t)
20 sec
s

20

0.75

Fig. 6.147 Input pulse of angular velocity

RJ + RM dv4 g Rb +
v =
dt +
v4 g

0.5

t, sec

RJ
RM dv4 g Rb
v =
+
+
+
v
dt
4 g

2

0.25

Flywheel
Rotational Inertia J

Electrical
Resistance R

Power Screw
Mass M

Voltage Source v(t)

Lubricating Film
Damping b

Fig. 6.150 Electromechanical system

6.8 Example Problems

393

DC Motor

Electrical
Resistance R

Flywheel
Rotational Inertia J

2
3

1
g

Power Screw
Mass M

v(t)

iM
g

3g= v4g

TM = i M
2
3

TM

J TS

FN

M b
g

Fig. 6.152 Linear graph of the system shown in Figs.6.150 and 6.151
Voltage Source v(t)

Lubricating Film
Damping b

Fig. 6.151 Electromechanical system, annotated to show nodes of


distinct values of the across variables, voltage, angular velocity, and
translational velocity

Rb + 2 2

RJ 2 + RM dv4 g
v =
+ v4 g
2 2
dt
Rb +

Rb + 2 2

RJ 2 + RM v
v =
+ [v]
2 2
Rb + t

200
100

v(t) 0
VDC

volt T t F volt F t
T F
+

i T
i
T v
i
v v

volt =
+ [v]

volt F
volt F
t


i v
i v

[v ] = + [v ] + [v ]
The units check.
Example Problem Five, Part B The system is de-energized
at time t = 0 . Solve the system equation for the voltage input
history given in Fig.6.153. Use Mathcad or MATLAB to
plot the velocity of the mass from time t = 0 to t = 6 sec.
Use superposition to construct input function v(t) from
scaled and time shifted Heaviside unit step functions,
Fig.6.154.

-100

t, sec

-200

Fig. 6.153 Input voltage pulse v(t)

200 VDC us(t)

200

volt
volt F T volt F

i v + i

T v

volt T t 2 volt F t
+

i v v + v
= i
[]
volt F T volt F t
i v + i

T v

200 VDC us(t-3)

100

v(t) 0
VDC
-100

t, sec

-200
-300

-400 VDC us(t-2)

-400
Fig. 6.154 Three scaled and time shifted step functions, which superpose to form the input pulse

Input Function:
v (t ) = 200VDCus (t ) 400VDCus (t 2) + 200VDCus (t 3)

Determine the unit step response of output variable v4g, assuming that the system is de-energized before application of
the unit step input. The output variable, the velocity of the
mass, v4g, and 3g, the angular velocity of the flywheel, are
dependent variables. The mass cannot move, if the flywheel
does not move. The value of one of the two variables determines the value of the other. Thus, either variable can be
used as the state variable of the system. For this problem,

6 Power Transmission, Transformation, and Conversion

394

5.0

it is convenient to use the velocity of the mass as the state


variable:

E sys = E J + E M
E sys =

1
J v3 g
2

E sys

1
Mv42g
2

v4g(t)

1
1
= J 32g + Mv42g
2
2

E sys =

1
J 2 + M v42g
2

( )

RJ + RM
Rb + 2 2
rad
+ (8 )( 20 kg )
m
2

Rb + 2 2

1VDC us ( ss )
=

RJ 2 + RM dv4 g ( ss )
+ v4 g ( ss )
dt
Rb + 2 2
0

N sec
N m
rad
(8 ) 2
+ 6.8
495

m
A
m

m
sec

(t 3)

0.52
+ 200 us (t 3) 2.97 10 4

1
e

sec

1VDC

(t 2 )

0.52
1
e
118.8 10 3

us (t 2)

sec

(t 3)

0.52
1
e
+ 59.4 103

us (t 3)

sec

6.8.6Example Problem 6: Hybrid Rotational


and Fluid System

Identify and formulate the unit step response


v4 g (0+ ) = 0 and v4 g ( ss ) = 2.97 10 4

m
0.52
1
e

sec

0.52
1
e

v4 g (t ) = 59.4 10 3

us (t )

sec

N m
rad

6.8
495

A
m
2

The response is a growth from zero of the form

(t 2 )

400 us (t 2) 2.97 10 4
1 e 0.52

sec

v4 g ( ss ) = 2.97 10 4

0.52
1
e

v4 g (t ) = 200 us (t ) 2.97 10 4

sec

v4 g ( ss ) =
1VDC
Rb + 2 2

v4 g ( ss ) =

= 0.52sec

The steady-state value of the unit step response

Form the response function by scaling and time shifting the


unit step response function with the same scaling and time
shifting as the input function (Fig.6.155):

t, sec

v4 g (t ) = 2.97 10 4

N sec
N m
rad
(8 ) 2
+ 6.8
495

m
A
m

v4 g (t ) = v4 g ( ss ) 1 e

RJ 2 + RM dv4 g
v=
+ v4 g
2 2
Rb +
Rb + 2 2 dt




(8 ) (3kg m 2 ) 495

Fig. 6.155 Response function v4g(t)

Determine the systems time constant and the steady-state


response of the output variable from the system equation:

-2.5

v4 g 0 = 0 = v4 g 0+

-5.0

The system is de-energized at time t = 0 . Hence

( )

cm
___
sec

2.5

m
sec

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 6in.
diameter pulley on the input shaft, a 5in. diameter pulley
on the shaft of a rotational inertia J = 1kg m 2, and a 4in.

6.8 Example Problems


Fig. 6.156 Belt-driven hybrid
rotational mechanical fluid
system

395

Inertia J
5 in. Dia.
Pulley

Fluid reservoir vented


to the atmosphere

Pump

Low pressure return

2 in. Dia. Idler

Hose, Fluid
Resistance R
High
pressure
discharge
4
in.
Dia.
2 in. Dia.
Pulley
Idler

Input Torque T(t)

Fig. 6.157 System schematic


annotated with nodes of distinct
values of the across variables,
angular velocity and pressure

6 in. Dia.
Pulley

Inertia J
5 in. Dia.
Pulley

Fluid reservoir vented


to the atmosphere

Pump

Low pressure return

2
g

1
3
Input Torque T(t)

diameter pulley on the shaft of a pump. The two other pulleys are idlers. The pump is rated at 15gallons per minute
at 200RPM. The pump pulls fluid from a reservoir vented
to the atmosphere, and discharges it to a hose with fluid resistance R = 100 MPa sec/m3. The flow returns to the reservoir.
Example Problem Six, Part A Use the linear graph method
to derive the system equation, which relates the input torque
to the pressure of the pumps discharge port, relative to
atmospheric pressure, and check its units.
Find the nodes of distinct values of the across variables,
angular velocity, and pressure on the systems schematic,
Fig.6.157.
Draw the linear graph, Fig.6.158. Draw the transducer
interfaces first, labeling the branches of the interfaces with
through variables, so that the branches are not subsequently
mistaken as other elements.

6 in. Dia.
Pulley

Hose, Fluid
Resistance R

High
pressure
discharge
4 in. Dia.
Pulley

Belt drive transformer equations The belts surface speed,


vs, is related to the angular speed of a pulley by the pulleys
radius:
vs = r1 1g

where r1 =

D1
2

D1
D
D
1g = vs = 2 2 g 2 g = 1 1g
2
2
D2
D
D1
1g = vs = 3 3 g
2
2

3g =

D1
1g
D3

Sum the power flows of two branches of each interface to determine the second transformer equation for each interface:

P + P = 0 T 1g + T 2 g = 0

6 Power Transmission, Transformation, and Conversion

396

2g = f 1g

3g = g 1g QP = h 3g

T(t) T

T J

T TP

QP R

200

rev 2 rad 1 min


rad
= 20.9
min 1 rev 60 sec
sec

The transducer ratio is


Qp

volume flow rate


=
=
angular velocity 3 g

g
Fig. 6.158 Linear graph of belt-driven system

Qp

3 g
T 1g + T

D
D1
1g = 0 T = 2 T
D1
D2

D
D1
1g = 0 T = 3 T
D1
D3

6in.
1g
5in.

2 g = 1.2 1g

5in.
T =
T
6in.

T = 0.833T

6in.
1g
4in.

3 g = 1.5 1g

4 in.
T =
T
6 in.

T = 0.667T

3g =

m3
rad

m3
3 g
rad

Qp = h 3 g

PA + PB = 0 TP 3 g + QP p4 g = 0
TP 3 g + h 3g p4 g = 0

TP = h p4 g

Energetic Equations
Continuity:
T (t ) = T + T

1g = 1g

D1 6 in.
=
= 1.2 and
D2 5 in.

2g = 2g

Elements: TJ = J

D1 6 in.
=
= 1.5
D3 4 in.

Pump transducer equations The pump is rated at 15gallons per minute at 200RPM. The volume must be converted
to cubic meters and the angular velocity to radians per second:
3
1min
15gal 3.79 liters 1m

min 1gal 1, 000 liters 60sec

= 9.48 10 4

T TJ = 0 T TP = 0 QP QR = 0

Compatibility:

It is easier and less error prone to write the transformer equations using single symbols in place of ratios or values:
f

Q p = 4.54 10 5

Sum the power flows to derive the second transducer equation

There is no need to convert the pulleys diameters from inches to meters, since the diameters only appear in ratios.

2g =

m3
rad

Define h 4.54 10 5

P + P = 0 T 1g + T 3 g = 0
T 1g + T

= 4.54 10 5

m3
sec
rad
20.9
sec

9.48 10 4

3g = 3g

d 1g

p4 g = p4 g

p4 g = RQR

dt

D1 6 in.
=
= 1.2
D2 5 in.

2 g = f 1g

T =

D1 6 in.
=
= 1.5
D3 4 in.

3 g = g 1g

1
T = T
g

h 4.54 10 5

m3
rad

Energy: E sys = E J

1
T
f

Q p = h 3 g TP = h p4 g

EJ =

1
J 22g
2

Reduction: Input T(t), Output p4g


m3
sec

T (t ) = T + T

T (t ) = f T + g T

6.8 Example Problems

397

T (t ) = f TJ + g TP T (t ) = f J
d 1g

T (t ) = f 2 J
T (t ) =

dt

d 2g
dt

T(t) 900
N m

h g p4 g

600

h g p4 g

300

f J d 3 g
h g p4 g
g
dt
2

-1

f 2 J dQP
h g p4 g
T (t ) =
g h dt
T (t ) =

f J dp4 g
h g p4 g
g h R dt

Tt
[ J ] =

t, sec

300 N m us (t+7)

p = RQ

300

-7

0 1

7 8
t, sec

T(t) -300
Nm

p
[ R] = Q

-600

600 N m us (t-2)

600

[ h] = = p

-900 N m u s(t-6)

-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
+ p4 g
g h T ( t ) = 2 2

g h R dt
Tt

1 J p
p p
h T = h 2 R t + [ p ] T T = 2 p t + [ p ]
h

Substitute in parameter values. Express MPa as Pa:

T t

p
p

T
=


+ [ p]
T Q T p t
p Q

[ p] = [ p] + [ p]

The units check:

1
f 2 J dp4 g
T (t ) = 2 2
+ p4 g
gh
g h R dt
2

Fig. 6.159 Input torque

Check consistency of the units in terms of the power variables and time:
d
T=J
dt

f 2 J dp4 g
1
+ p4 g
T (t ) = 2 2
gh
g h R dt

D1
D J
2

dp4 g
1
T (t ) =
+ p4 g
2
D1
dt
D1
2
h
D3
D ( h ) R
3

1
T (t )
3
6in.
5 m
4in. 4.54 10 rad
2

6in.
2
5in. 1kg m
2

dp4 g

dt
6in.
6 Pa sec
5 m
4in. 4.54 10 rad 100 10 m3

14.7

+ p4 g

dp4 g
kPa
+ p4 g System Eq. in form
T (t ) = 3.11
Nm
dt

Example Problem Six, Part BThe system is running in steady-state under a previously applied torque of
300 N m , when the applied torque is increased to 900 N m
at time t = 2, and then decreased to zero at time t = 6, as
shown in Fig.6.159. Solve the system equation for the input
torque history given and plot the pumps discharge pressure
from time t = 0 to t = 6 + 6 sec.
Create the input torque function by superposing three
scaled and shifted step inputs, Fig.6.160:

6 Power Transmission, Transformation, and Conversion

398

T (t ) = (300 N m ) us (t + 6 )

+ ( 600 N m ) us (t 2) (900 N m ) us (t 6)

Calculate the unit step response of the system from its deenergized state. From the system equation find the time
constant and the steady-state response to a unit step input,
T (t ) = 1 N m us (t ).
1
f 2 J dp4 g

T (t ) = 2 2
+ p4 g
gh
g h R dt

p4g(t)

MPa

-2
-4
-6
-8

-10
-12
-20

-10

6in
2
5in 1kg m

The unit step response is a stable exponential growth from


zero of the form

p4 g (t ) = p4 g ( ss ) 1 e

3
6in
5 m
6 Pa sec
4in 4.54 10 rad 100 10 m3

g 2 h2 R

2
t
1
p4 g ( t ) =
T0 1 e f J
g h

= 3.11sec

1
f 2 J dp4 g ( ss )
1N m us ( ss ) = 2 2
+ p4 g ( ss )
1
dt
gh
g h R
0
1
p4 g ( ss ) =
1N m
gh
p4 g ( ss ) =

m3
(1.5) 4.54 105
rad

1N m

u.s.

p4 g

( )

3g

( )

Rhg
2 g 0+
f

( )

Create the response function, Fig.6.161, by scaling and


timeshifting the unit step response function with the same
scaling and time shifting as the input torque function, T(t):

p4 g (t ) = 300 ( 14.7 kPa ) 1 e 0.311(t + 6 ) us (t + 6 )

+ 600 ( 14.7 kPa ) 1 e 0.311(t 2) us (t 2)

900 ( 14.7 kPa ) 1 e

0.311(t 6)

) u (t 6)
s

p4 g (t ) = 4.41MPa 1 e 0.311(t + 6 ) us (t + 6 )

8.82 MPa 1 e 0.311(t 2) us (t 2)

+ (13.2 MPa ) 1 e 0.311(t 6) us (t 6)

( )

p4 g 0+ = R h g 1g 0+
p4 g 0 + =

u.s.

( ) p (0 ) = RQ (0 )
(0 ) = R h (0 )
4g

3.11
1

m3

(1.5) 4.54 105


rad

1 Nm

p4 g (t ) = 14.7 kPa 1 e 0.311t

Determine the initial value of output variable, p4 g (0+ ) , in


response to a unit step input. The system is assumed to deenergized at time t = 0 . The state variable is the angular
velocity of the rotational inertia 2g. The value of the state
variable remains unchanged during the transition of the step
input, 2 g (0 ) = 0 = 2 g (0+ ). Calculate the value of output
variable p4 g (0+ ) using the values of the state variable and
the unit input torque at time t = 0+ .

( )

Substitute in parameter values and evaluate the constant


terms of the unit step response
p4 g (t ) =

p4 g ( ss ) = 1, 200 Pa = 1.2 kPa

p4 g 0+ = RQR 0+

20

Fig. 6.161 Response function p4g(t). The transducer signs, which led
to negative pressure, would be fixed before these results were used

f2J
= 2 2
g h R
2

10

t, sec

( )

p4 g 0+ = 0

Summary
Fig. 6.162 Linear graph
transformer and transducer sign
conventions for positive power
flow

399

b
A + B = 0
PowerA

A = TA 1g

Power B

Transformer
or
Transducer

B = TB 2g

N
N1

2
1g = - ___
2g

PowerA

Power B
TB

TA

Summary
Transformers and transducers interface subsystems with a
larger dynamic system. Transformers interface like subsystems, and transducers interface dissimilar subsystems. Power
flows across the interface between the subsystems, but there
is no direct connection between the through variables in the
two branches of the transducer or transformer interface. The
power flowing down one branch of the interface, equal to the
product of the through variable and the across variable difference of that branch, is equal to the product of the through and
across variables in the other branch.
The sign convention is based on (a) assuming that the
power flows on both sides of the interface is positive and (b)
orienting the through variables in the two branches of the
interface towards ground. Because the positive direction of
power flow is into the transformer or transducer interface,
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
on one side of the interface to a power variable on the other
side, will have a negative sign. Although a negative sign
can be physically possible, as in the case of a pair of gears,
often, the negative sign is nonsense. The negative sign is
a result of the linear graph sign convention. When there is
a non-sensical negative sign, it should be tolerated, until
the derivation of the system equation is completed. Then
fix the sign, if necessary, before using the results. Correcting the sign prior to completion of the reduction leads to
further errors. Sign conflicts between subsystems are common, because the subdisciplines of mechanical engineering
have different histories.
There are three equations associated with a transformer
or transducer. The first equation is the summation of power
flow into the interface, Fig.6.162a, which is the sum of the
products of the power variables in the two interface branches. The second and third equations each relate a single power
variable on one side of the interface with a single power variable on the other side. The latter two equations are useful

in the derivation of a system equation, since there is a onefor-one power variable substitution. The power summation
equation is valuable but inefficient in a reduction, because
the power variable substitution is one-for-three.
It is often possible to derive a relationship between the
displacements or velocities on the two sides of the interface,
using the geometry of the mechanism of the volume displacement of a pump or hydraulic motor. In other cases, such
as with electric motors, generators, or turbines, experimental
data is needed to establish a transducer equation. Once a relationship is obtained, then substitution into the power summation yields the second one-to-one equation, for example,

1g =

N2
2g
N1

PA + PB = 0

1g TA + 2 g TB = 0

N2
2 g TA + 2 g TB = 0
N1

TB =

N2
TA
N1

When drawing linear graphs of systems with transformers


and transducers:
1. Find nodes of distinct values of the across variables for
each subsystem, and identify those nodes on the schematic.
2. Draw and identify the nodes of the linear graph.
3. Draw the two interface branches for each transducer or
transformer, and name the through variable of each interface branch.
4. Add the remaining elements of the system between their
respective nodes.
5. There is no branch across a transformer or transducer interface! There is no through variable flow across the interface. Power flows across the interface.
A transformer constant is dimensionless. A transducer constant has two sets of units. Generally, both sets must be used
to check the units of a system equation with a transducer.

6 Power Transmission, Transformation, and Conversion

400
Fig. P6.1 DC motor driving an
inertial load, JL

Flywheel
Rotational Inertia JL

Electric Motor
Transduction KT
Rotational Inertia J M
Electrical Resistance R

Bearings
Damping b

Voltage Source v(t)

Fig. P6.2 Hybrid fluid-translational mechanical system

Fluid Power Unit


Pressure Source p(t)
Breather Cap (vent to atmosphere)
Hydraulic Line
Resistance R 1
Fluid Accumulator
Capacitance C

Hydraulic Piston-Cylinder
Piston diameter D
Damping b between Piston
and Cylinder

Hydraulic Line
Resistance R2

Press Die
Mass M

Hydraulic Line
Resistance R 3
Inertance I

Problems
Problem 6.1 Draw the linear graph for the system shown in
Fig.P6.1 and derive the transformer or transducer constants
equations.
Problem 6.2 Draw the linear graph for the system shown in
Fig.P6.2 and derive the transformer or transducer constants
equations.
Problem 6.3 Draw the linear graph for the system shown in
Fig.P6.3 and derive the transformer or transducer constants
equations.
Problem 6.4 Draw the linear graph for the system shown in
Fig.P6.4 and derive the transformer or transducer constants
equations.
Problem 6.5 Draw the linear graph for the system shown in
Fig.P6.5 and derive the transformer or transducer constants
equations.

Problem 6.6 Draw the linear graph for the system shown in
Fig.P6.6 and derive the transformer or transducer constants
equations.
Problem 6.7 Draw the linear graph for the system shown in
Fig.P6.7 and derive the transformer or transducer constants
equations.
Problem 6.8 Draw the linear graph for the system shown in
Fig.P6.8 and derive the transformer or transducer constants
equations.
Problem 6.9A rotational mechanical system is shown in
the Fig.P6.9. An angular velocity source, (t), acts on the
input shaft of a fluid coupling. The output shaft of the fluid
coupling drives a pinion with N1 teeth, which is engaged in
a rack with mass M and N2 teeth per meter. The gear has
N3 teeth and inertia J2. The bearings on the input shaft collectively have damping b1. The bearings on the output shaft
collectively have damping b2.

Problems
Fig. P6.3 Rotational mechanical
system

401

Compliant Shaft
Spring Constant K

Pulley
Diameter D1

Flywheel
Rotational Inertia J1
Flywheel
Rotational Inertia J 2

Pulleys Diameters
D2 and D3

Angular Velocity
Input (t)

Fluid Coupling
Damping b

Pulley
Diameter D4
Fig. P6.4 Hybrid fluid-translational mechanical system

Fluid Power Unit


modeled as a
Pressure Source
Hydraulic Line
Fluid Resistance R

Dashpot
Damping b
L1
L2

Tank vented to
the atmosphere
Pivot
Hydraulic Cylinder
Diameter D
Linkage
Rotational Inertia J
Fig. P6.5 Rotational mechanical
system

Rotational
Inertia J

Spring K

Bearing
Damping b 3

Compliant Shaft
Torsional Spring K 2
N4
N2

Compliant Shaft
Torsional Spring K 1
Fluid Coupling
Damping b1
Angular Velocity
Source, (t)

N3
N1

Bearing
Damping b 2

6 Power Transmission, Transformation, and Conversion

402
Fig. P6.6 Electromechanical
system

Flywheel
Rotational Inertia J L

Electric Motor
Transduction K T
Rotational Inertia J M

Gears 2 and 3 are keyed


to the same shaft
N2

N4

Voltage Source v(t)

N1
N3
Bearings
Damping b

Electrical Resistance R

Fig. P6.7 Translational mechanical system

Lubricating Film
Damping b 2

Mass M
Spring K2

Spring K 1

Pivot

Dashpot b1

L3
L2

Force Source F(t)

Fig. P6.8 Electromechanical


system. The cross on the top
linkage is the centerline of the
solenoids magnetic force

L1

L1

L2

Top Linkage
Rotational Inertia J

Solenoid
Coil Resistance R
Motor Constant
Voltage Source

Dashpot
Damping b
Spring K

Frame

6.9.a 
Draw the linear graph of the existing system and
determine the transformer equations.
6.9.b Draw the equivalent linear graph, if the gear set is
eliminated, and inertias J1 and J2, the bearings b2, and
the torsional shaft with spring constant K are replaced
by equivalent elements attached to the input shaft.
6.9.c 
Calculate the equivalent elemental parameters for the
equivalent system of part b.

Problem 6.10 A mechanical system consisting of a torque


source and a rack and pinion is shown in Fig.P6.10. The
pinion has N1 = 8 teeth. The rack has N 2 = 4 teeth/in. The
torque source acts on a pinion with mass moment of inertia
J = 0.05 kg m 2 . The pinion is engaged with a rack of mass
M = 1 kg. The rack slides on a lubricating film with damping
b = 6 N sec/m .

Problems

403

Fig. P6.9 Hybrid rotationaltranslational mechanical system

Pinion
N1 teeth

Inertia J

Angular Velocity
Source, (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 rotationaltranslational mechanical system

Pinion
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:


i. The velocity of the rack.
ii. The force acting to accelerate the rack.
iii. The force acting to shear the lubricating fluid
film.
Check their units.
The system is at rest at time t = 0 , when a torque
T (t ) = 60N m us (t ) is applied.
6.10.b Solve the system equations.
6.10.c Plot the responses from t = 0 to t = 6 using Mathcad
or MATLAB.
Problem 6.11 An electromechanical schematic of a DC
motor is shown in Fig.P6.11a. The motors resistance is
R = 4 . The relationship between the motor current and the

motor torque is TM = KT iM , where KT = 8 N m/A . The motors mass moment of inertia is J M = 0.3 kg m 2 . The motor
turns a flywheel with mass moment of inertia J L = 2 kg m 2
and damping b = 0.1 N m sec/rad .
6.11.a Derive the system equations for
i. The current through resistance R.
ii The voltage drop across resistance R.
iii. The back EMF.
iv. The motors torque TM.
v. The torque acting through damping b.
vi. The torque acting to accelerate mass moment of
inertia JM+JL.
vii. The angular velocity of mass moment of inertia
JL .
and check their units.

6 Power Transmission, Transformation, and Conversion

404
Fig. P6.11a Schematic of a DC
motor. b Voltage v(t) applied to
the motor

Electric Motor
Transduction K T
Inertia J M

Flywheel
Inertia JL

100

Resistance R

v (t)
75
VDC
50

Voltage
Source
v(t)

Bearings
Damping b

t, sec
0.5

1.0

-25
-50

Fig. P6.12a Translational


mechanical system. b Force input
F(t)

Lever
Pivot

LA

F(t),N

40
30
20

F(t)

LB

10

LC

-10

The motor was running in steady-state at time t = 0 under


the previously applied step input, when the pulse shown in
Fig.P6.11b was applied.
6.11.b Solve the system equations.
6.11.c Plot the responses from time t = 0 to t = 6 sec
using Mathcad or MATLAB.
Problem 6.12 A mechanical system is shown in Fig.P6.12a.
A force source acts on a lever with lengths LA = 1 m,
LB = 2 m, and LC = 1 m. Connected to the lever are a spring,
K = 80 N/m, and a damper, b = 4 N sec/m .
6.12.a Derive the system equations for the following:
i. The force acting through the spring.
ii. The force acting through the damper.
iii. The velocity of the point of application of the
input force.
Check their units.
The system is in steady-state under the action of a previously applied step input of 20N, before the force input, F(t),
plotted in Fig.P6.12b, is applied.
6.12.b Solve the system equations.
6.12.c Plot the responses from time t = 0 to t = 2 sec + 6
using Mathcad or MATLAB.

0.5 1.0 1.5 2.0

2.5

t, sec

Problem 6.13 The rotational mechanical system shown in


Fig.P6.13a consists of a velocity source acting on the input
shaft of rotational damper b1 = 100 N m sec (a drag cup).
The output shaft of damper b1 drives a pinion, N1 = 12 teeth
meshed with two gears with N 2 = 24 and N 3 = 36 teeth.
The gears, in turn, drive the input shaft of rotational damper
b2 = 100 N m sec and rotational inertia J = 25 kg m 2 .
6.13.a Derive the system equations for the following:
i. The torque from the velocity source.
ii. The torque acting to accelerate mass moment of
inertia J.
iii. The angular velocity of mass moment of inertia J.
iv. The torque acting through damper b2.
Check their units.
The system was in steady-state under a previous step
input of (t)=5rad/sec before the angular velocity input,
plotted in Fig.P6.13a, was applied.
6.13.b Solve the system equations.
6.13.c Plot the responses from t = 0 to t = 1 + 6 sec using
Mathcad or MATLAB.
Problem 6.14 A translational mechanical system is shown
schematically in Fig.P6.14. The system consists of a

Problems
Fig. P6.13a Rotational system.
b Angular velocity input

405

b
(t)
rad 10
___
sec

Shaft welded to plate


Damper b2
Gear
N2 teeth

Shaft rotates in
ideal bearing

Pinion
N1 teeth

0.5
-5

t, sec

(t)

Inertia J
Gear
N3 teeth

Fig. P6.14 Translational mechanical system

1.0

Damper b1

Velocity source.
Positive rotation direction
defined by the right-hand rule

Lubricating Film
Damping b2

Dashpot b1

Pivot

Mass M

Spring K

L3
L2

Force Source F(t)

force source F (t ) acting on a pivoted beam with lengths


L1 = 36in., L 2 = 48in., and L 2 = 36 in. Attached to the
beam are damper b1 = 200 N sec/m , spring K = 4, 000 N/m,
and mass M = 20 kg . There is a lubricating film under mass
M with damping b2 = 10 N sec/m .
6.14.a Derive the system equations for the following:
i. The force acting to accelerate mass M.
ii. The force acting through the damper b1.
iii. The force acting through the damper b2.
iv. The velocity of mass M.
v. The velocity of the point of application of the
input force.
Check their units.

L1

The system was de-energized for time t < 0, before the


input force F(t)=1,000Nus (t) acted on the system.
6.14.b Solve the system equations.
6.14.c Plot the responses from time t = 0 to t = 10 + 6
using Mathcad or MATLAB.
Problem 6.15 A model of an engine lathe is shown in
Fig.P6.15a. The motor and gear box is modeled as a torque
source driving a rotational inertia J = 2.5kg m 2. The motor
drives a lead (or power) screw, which is supported by bearings in the tail stock. The lead screw is modeled as having
negligible torsional compliance. The lead screw bearings
have damping b1 = 10 N m sec/rad . The lead screw pitch
is 2 threads per inch. The lead screw engages a lead nut

6 Power Transmission, Transformation, and Conversion

406
Fig. P6.15a Schematic model
of an engine lathe. b Torque
input T(t)

Load Carriage

Lead
Screw

Lead
Nut

Motor
Inertia

Lead Screw
Bearings with
Damping b1

b
200

T(t)
150
N m

T(t) J

100
50
0

Lathe Way

Fig. P.6.16 Hybrid rotationaltranslational mechanical system

Lubricating Film
Damping b2

Angular Velocity Source, (t)


Compliant Shaft
Torsional Spring K

Rack, N2 teeth
per inch
Mass M

t, sec

b
(t)
rad
___
sec

Pinion, N1 teeth

80
60
40
20

-20

0.5 1.0 1.5 2.0 2.5

t, sec

Lubricant Film
Damping b

attached to the load carriage. The lead nut moves the load
carriage with mass M = 150 kg along the lubricated ways of
the lathe. The lubricating film between the load carriage and
the lathes ways has a damping coefficient b2 = 60 N sec/m
6.15.a Derive the system equations for the following:
i. The torque acting to accelerate the motors mass
moment of inertia J.
ii. The angular velocity of the lead screw.
iii. The force acting to accelerate the load carriage
mass M.
iv. The velocity of the load carriage mass M.
Check their units.
The system was at rest before the input torque T(t), plotted in Fig.P6.15b, was applied.
6.15.b Solve the system equations.
6.15.c Plot the responses from t = 0 to t = 3 + 6 using
Mathcad or MATLAB.

spring constant K = 100 N m sec/rad . The shaft drives a


pinion with N1 = 10 teeth. The pinion is engaged in a rack.
The rack has N 2 = 4 teeth/in . The rack is attached to mass
M = 10 kg. The rack and mass slide on a lubricating film with
damping b = 20 N sec/m.
6.16.a Derive the system equations for the following:
i. The velocity of the rack.
ii. The force acting to accelerate the rack.
iii. The force acting to shear the lubricating fluid
film.
Check their units.
The system is in steady-state under a step input of angular
velocity of 40rad/sec and time t = 0 , when the angular
velocity pulse shown in Fig.P6.16b is applied.
6.16.b Solve the system equations.
6.16.c Plot the responses from t = 0 to t = 2 + 6 using
Mathcad or MATLAB.

Problem 6.16 A mechanical system is shown in Fig.P6.16a.


The angular velocity source drives a compliant shaft with

Problem 6.17 A fluid translational mechanical system is shown in Fig.P6.17a. The system consists of a
pump modeled as a pressure source, two fluid resistances

Problems

407

Fig. P6.17a Fluid translational


mechanical system. b Pressure
input

a
C
p(t)

R1
R2

Hydraulic
Piston
Area A

5,000

p(t) 4,000
psi
3,000

1,000
0

Fig. P6.18a Translational


mechanical system. b Input
force F(t)

400

Dashpot b 1

100

Spring K1

L1
Dashpot b 2
L2
L3

Spring K 2

R1 = 10 MPa sec/m3 and R2 = 100 MPa sec/m3, an accumulator (fluid capacitor) C=5.5L/MPa, an hydraulic cylinder with a piston of diameter D = 6 in. and mass M = 2 kg .
6.17.a Derive the system equations for the following:
i. The volume flow rate from the pump.
ii. The force acting to accelerate the piston.
iii. The velocity of the piston.
Check their units.
The system was in steady-state under a previous step
input of 2,000psi, when the input plotted in Fig.P6.17b was
applied at time t = 0 .
6.17.b Solve the system equations.
6.17.c Plot the responses from t = 0 to t = 6 using
Mathcad or MATLAB.
Problem 6.18 A translational system consisting of a force
source F(t), a lever, two springs, and two dashpots is shown
in Fig.P6.18a. The parameter values are tabulated.
b2

K1

K2

N sec
125
m

N
175
m

200

300
200

Pivot

N sec
48in. 60in. 36in. 100
m

2.0

t, sec

Force Source F(t)

Table P6.18 Parameter values


L1
L2
L3
b1

1.0

N
m

F(t)
0
N
-100

0.5 1.0 1.5 2.0

t, sec

-200
-300
-400

6.18.a Derive the system equations for the following:


i. The force acting through spring K1.
ii. The force acting through spring K2.
iii. The force acting through the damper b1.
iv. The force acting through the damper b2.
v. The velocity of the point of application of the
input force.
Check their units.
The system is at de-energized at time t = 0 when the system is subjected to input force F(t), plotted in Fig.P6.18b.
6.18.b Solve the system equations.
6.18.c Plot the responses from time t = 0 to t = 2 + 6
using Mathcad or MATLAB.
Problem 6.19 A rotational mechanical system is shown
schematically in Fig.P6.19. The system is driven by an angular velocity source which is connected to the input shaft,
by a fluid coupling with damping b1. The output shaft of the
fluid coupling connects to a compliant shaft with spring constant K, which turns the pinion of a gear set. The pinion has
N1 teeth. Gear N2 drives a shaft which rotates mass moment
of inertia J and is supported by a bearing with damping b2.

6 Power Transmission, Transformation, and Conversion

408
Fig. P6.19 Rotational mechanical system

Bearing Damping b 2
Pinion
N1 teeth

Inertia J

Compliant Shaft
Spring Constant K
Fluid Coupling
Damping b1

Gear
N 2 teeth

(t)
Angular velocity source

DC Motor
Transduction KT

Fig. P6.20 DC motor driving an


inertial load through a compliant
shaft

Bearings Damping b1

Voltage Source
v(t)

Compliant Shaft
Spring Constant K

Resistance R

Rotational Inertia J
Bearings Damping b2
Table P6.19 Parameter values
b1
K
N1
80

N m sec
Nm
1,000
rad
rad

12

N2
36

b2

5 kg m 2

0.2

N m sec
rad

6.19.a Derive the system equations for the following:


i. The torque acting through spring K.
ii. The angular velocity of inertia J.
iii. The velocity difference across spring K.
Check their units.
The system is at de-energized at time t = 0 when the system is subjected to the step input torque of 800N-m.
6.19.b Solve the system equations.
6.19.c Plot the responses from using Mathcad or MATLAB.
Problem 6.20 An electromechanical schematic of a DC motor
is shown in Fig.P6.20. The motor is powered by a voltage
source v(t). The motors resistance is R = 4 . The motors

bearings and brushes have damping b1 = 0.4 N m rad/sec.


The relationship between the motor current and the motor
torque is TM = KT iM , where KT = 8 N m/A . The motor turns
a compliant shaft with spring constant K = 500 N m/rad.
The shaft turns a flywheel with mass moment of inertia
J = 2 kg m 2 and damping b2 = 0.1 N m sec/rad .
6.20.a Derive the system equations for the following:
i. The current through resistance R.
ii. The back EMF.
iii. The motors torque TM.
iv. The torque acting through the shaft.
v. The angular velocity of mass moment of inertia J.
Check their units.
The motor de-energized when a step input
v(t ) = 48VDC us (t ) was applied.
6.20.b Solve the system equations.
6.20.c Plot the responses from time t = 0 to t = 6 sec
using Mathcad or MATLAB.

References and Suggested Reading

References and Suggested Reading


Budynas RG, Nisbett KJ (2011) Shigleys mechanical engineering
Design, 9thedn. McGraw-Hill, New York
Fitzgerald AE etal (2003) Electric machinery, 6thedn. McGraw-Hill,
New York

409
Ogata K (2003) System dynamics, 4th edn. Prentice-Hall, Englewood
Cliffs
Rowell D, Wormley DN (1997) System dynamics: an Introduction.
Prentice- Hall, Upper Saddle River
Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
dynamics. Addison-Wesley, Reading

Vector-Matrix Algebra and the State-Space


Representation of Dynamic Systems

Abstract

An alternative to deriving a high-order differential system equation is to represent the system as a set of simultaneous first-order differential equations of the systems 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 control designer to arbitrarily tailor the response of a system, within the physical limitations of
the hardware.

7.1Overview
Up to this point, we have derived first- and second-order system equations and solved them, using the method of undetermined coefficients and Laplace transformations. However,
many machines and processes have more than two independent energy storage variables and, consequently, are higherorder systems. The effort to reduce the equation list to a differential system equation increases, roughly with the square
of the number of the branches in the linear graph. The reduction of a third-order system is roughly nine times the effort of
the reduction of a first-order system. We do have an alternative. The equation list does not need to be reduced to a single,
higher-order differential equation. We can formulate a set of
simultaneous, first-order differential equations containing
the information needed to describe the system. The information needed to describe the state of the system at any
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
variable. The energy storage variables of a system are called
the state variables, since they allow us, with knowledge of
the input, to calculate all the power variables in the system.
The set of simultaneous, first-order differential equations,
written in terms of the state variables and time, are called the

state equations. The set of simultaneous, first-order differential state equations must be solved numerically, to yield
the energy storage (or state) variables. If our output variable
is one of the energy storage variables, then we are done, once
we have solved the state equations. Often, though, we are
interested in the force in a damper or other non-energy storage variable. In those cases, a two-step process is used, in
which a set of simultaneous algebraic output equations,
formulated in terms of the state variables, is solved, once the
state variables have been determined.
We will use Mathcad to solve this set of simultaneous differential (state) equations for the state variables, from which
we can calculate every power variable in the system. In addition to easing the reduction of higher-order systems, the
numerical solution of the system, formulated using the statevariable method, frees us from the restriction of using linear
elemental equations, imposed by our analytical method.

7.2 Vector-Matrix Algebra


Matrix notation was invented for the manipulation of systems of simultaneous equations. We will use matrices to express the state equations and the output equations in a simple
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

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

412

to manipulate mathematical symbols. There are many different types of algebra. The collection of rules for manipulating matrices is vector-matrix algebra, or, more simply, just
matrix algebra, because a vector is a matrix with only one
row or column. Linear algebra is a set of rules that defines
the allowable operations on systems of linear simultaneous
equations represented as matrices. The easiest way to understand the manipulations of matrix algebra or linear algebra
is to work with a system of equations, consisting of either
two or three simultaneous equations. The rules that apply to
these systems of equations can be extended to any system of
equations.
Before we begin, we must address a bit of confusing terminology. As you know, a vector is a physical quantity
with magnitude and direction, such as a force vector. Possessing magnitude and direction are necessarybut not sufficient characteristics. To be a vector, the physical quantity
must also obey the rules of vector algebra. For example, a
force vector can be decomposed into its components in orthogonal directions
F = Fx + Fy + Fz = Fx
u x + Fy
u y + Fz
uz
which can be expressed as the product of two vectors

Fx
u x + Fy
u y + Fz
u z = Fx

Fy

ux

Fz
uy

u z

where the magnitudes of both the force components and the


unit vectors in the principal directions are represented as ordered lists. We often use a further simplified form by omitting the unit vectors in the principal directions, and presenting the magnitudes of the force components, as Cartesian
coordinates in an ordered list.
Fx

Fy

Fz

It is this form which creates the confusion in terminology.


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
as a vector, as it represents a physical quantity with magnitude 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 twoor three-dimensional space. Most of us base our conception
of a vector on position vectors, but are they truly vectors?
Do they obey the rules of vector manipulation? They do, depending on the meaning you apply to the position. A vector

should retain its properties as it is slid around in space,


allowing us to perform vector addition using the trapezoidal
construction. Does a position vector retain its meaning, when
its tail is no longer at the origin? It does, if you are willing to
interpret position as displacement. We will call a position vector a vector. We will also refer to complex numbers
as vectors because they obey most of the rules of vector manipulation, although they do not multiply as vectors. Complex numbers are clearly a special case. They do not exist in
three dimensions. Fourth-dimensional complex numbers are
quaternions invented by Hamilton.
Unfortunately, the term vector has been extended from
an ordered list of coordinates to any ordered list, primarily
by the use of vector in computer science. For example,
we now speak of data vectors, which are, in fact, time series, or sequences of scalar variables, which have none of
the physical or mathematical properties of true vectors, such
as force vectors. It is important to be aware that there are
multiple meanings. The mathematics which can be applied
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
with magnitude and direction, and other quantities which are
only ordered lists, referring to the latter as a column matrix
or a row matrix, rather than the alternatives, column vector and row vector.

7.2.1 Matrix Addition


Matrix addition and subtraction are defined only for vectors
and matrices with the same number of rows and columns.
Addition and subtraction are performed element by element.
For example, given the matrices
a 1,1
A
a 2,1

a 1,2
and
a 2,2

h1,1
H
h 2,1

h1,2
h 2,2

then

a 1,1
A+H=
a 2,1

a 1,2
+
a 2,2

h1,1

h 2,1

h1,2 a 1,1 + h1,1


=
h 2,2 a 2,1 + h 2,1

a 1,2 + h1,2
a 2,2 + h 2,2

(7.1)
and


a 1,1 a 1,2
AH=

a 2,1 a 2,2

h1,1 h1,2 a 1,1 h1,1 a 1,2 h1,2


h
=

2,1 h 2,2 a 2,1 h 2,1 a 2,2 h 2,2

(7.2)

7.2 Vector-Matrix Algebra

413

7.2.2 Matrix Multiplication


The matrix representation of simultaneous equations and
the procedure for matrix multiplication are both illustrated
by expressing a system of simultaneous equations in matrix
form. Consider the system of equations


y1 = a1,1 x1 + a1,2 x2
y2 = a 2,1 x1 + a 2,2 x2

(7.3)

where xc and yr are variables, ar,c are constant coefficients,


and subscripts r and c indicate row and column, respectively.
This system of equations can be written in matrix notation as


y1 a1,1
y = a
2 2,1

a1,2 x1
a 2,2 x2

Generalizing, the result of matrix multiplication of matrices


P and Q will have as many rows as P and as many columns
as Q. It is important since matrix multiplication is evaluated as row times column, first element times first element,
second element times second element, etc., the number of
columns in P must equal the number of rows in Q or the multiplication PQ cannot be evaluated.
Having defined matrix multiplication in expanded
form, we can now simplify the notation and express the vector-matrix equation vectors, y and x, and the matrix, A, as
variables


(7.4)

a1,1 a1,2
x1
y
where 1 and are column matrixes and

x2
a 2,1 a 2,2
y2
is a two by two matrix, since it has two rows and two columns.
Matrix multiplication is performed by multiplying a row
by a column, element by element, and then summing the
products. The mnemonic is R C.
For example, to multiply

a1,1
a
2,1

a1,2 x1
a 2,2 x2

begin by multiplying the first row of the coefficient matrix A


and the column matrix x
a1,1

x1
a1,2 = a1,1 x1 + a1,2 x2
x2

x1
a 2,2 = a 2,1 x1 + a 2,2 x2
x2

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.
a1,1
a
2,1

a1, 2 x1 a1,1 x1 + a1, 2 x2


=
a 2, 2 x2 a 2,1 x1 + a 2, 2 x2

(7.5)

a1,1
A
a 2,1

a1,2
a 2,2

(7.6)

x1
x
x2

(7.7)

where bold is used to denote a vector or matrix. Conventionally, vectors are denoted by lower case, and matrices by
upper case, where possible. This allows us to write the system of equations, Eqs.7.1 and 7.2, in a more compact but
abstract form


y = Ax

(7.8)

Although compact and convenient to manipulate, a matrix


equation written using single variables, such as y = Ax, reveals little about the system of equations it represents. For
example, until vectors y and x, and matrix A are defined, the
equation y = Ax could represent a system of three simultaneous equations


next, multiply the second row of the matrix A and the column matrix x
a 2,1

y1
y
y2

y1 a1,1 a1,2
y = a
2 2,1 a 2,2
y3 a 3,1 a 3,2

a1,3 x1

a 2,3 x2
a 3,3 x3

(7.9)

or 300,000 as in the case of large finite element models




a1,1 a1,2
y1
a
y
a 2,2
2 = 2,1


a
a
y
n ,2
n
n ,1

a1, n
a 2, n

an , n

x1
x
2


xn

(7.10)

It is frequently necessary to expand a matrix equation written in the form of Eq.7.6 to that of Eq.7.3, in order to confirm the matrix algebra. Whenever you are in doubt as to the
validity of your algebra or the meaning of a matrix-algebra

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

414

equation, expand the equation to a system of simultaneous


equations to evaluate it.
Matrix multiplication is not commutative!
The most common errors in matrix algebra are (1) attempting to perform an operation using two terms, which
have incompatible numbers of rows and columns and (2)
forgetting that matrix multiplication is not commutative.
For example, given the row matrix u and the column matrix v
u = u1 u 2 and
then


v1
v=
v2

v1
uv = u1 u 2 = u1v1 + u 2 v2
v2

7.3Operating on a Vector-Matrix Expression


with a Linear Operator
An operator is a linear operator, if doubling the input to the
operator doubles the output. Differentiation with respect to
time is a linear operator. For example,


v1u1
v1
vu = u1 u 2 =
v2
v2 u1

v1u 2
v2 u 2

(7.11)

(7.12)

Clearly, uv vu. These two products demonstrate that the


number of rows and columns in the product of a matrix multiplication equals the number of rows in the first term of the
product, and the number of columns in the second term.

7.2.2.1 Pre- and Post-Multiplication


The fact that vector-matrix is not commutative makes the
position of a factor in a product important. The terms premultiplication and post-multiplication are used to distinguish product uv from vu. In the former, u pre-multiplies v.
In the latter, v pre-multiplies u. Alternatively, in the former, v
post-multiplies u and in the latter, u post-multiplies v.
7.2.2.2 Multiplication of a Matrix by a Scalar
The product of a vector or matrix and a scalar (or a single
valued function) is evaluated by multiplying each term in the
vector or matrix by the scalar. For example, given a vector
b1
B=
b 2

dy1 d
dy1 d
d
= (a1,1 x1 + a1, 2 x2 )
= (a1,1 x1 ) + (a1, 2 x2 )
dt
dt
dt
dt
dt

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

b1
b1c
Bc = c =
b
2
b 2 c

dy1
dx
dx
= a1,1 1 + a1, 2 2
dt
dt
dt
dy2
dx1
dx2
= a 2,1
+ a 2, 2
dt
dt
dt

(7.16)

This is expressed in matrix notation as




d y1 d a1,1
=

dt y2 dt a 2,1

a1,2 x1
a 2,2 x2

(7.17)

A constant matrix is a matrix whose elements are constants, such as the constants ar,c in Eq.7.15. Constant matrices are factored out of the derivative of matrix products,
just as scalar constants are factored out of the derivatives of
scalar products. Hence



(7.13)

The multiplication of a vector or matrix by a scalar is commutative


b1
b1c cb1
b1
Bc = c = = = c = cB
b 2
b2 c cb2
b 2

d y1 a1,1
=
dt y2 a 2,1

a1,2 d x1
a 2,2 dt x2

(7.18)

The same rule applies in the vector-matrix notation

and scalar variable c, the product is




(7.15)

Operating on a matrix or a matrix equation is analogous to


multiplying by a scalar. Differentiating the system of equations of Eq.7.1 with respect to time

which is a scalar. But




d 2x
dx
=2
dt
dt

(7.14)

dy dAx
=
dt
dt

(7.19)

where vectors x and y are variables and matrix A is constant, resulting in




dy
dx
=A
dt
dt

(7.20)

7.4 Transpose of a Matrix

415

7.3.1Laplace Transformation of Matrix or a


Vector-Matrix Expression

yields


An important linear operator is the Laplace Transform.

and

L { f (t )} = f (t ) e

st

dt F ( s )

(7.21)

The functions f(t) and F(s) are functions of time t and the
complex variable s, respectively. The complex variable
s = + jw in the exponential within the Laplace integral is
the same complex variable s in the trial solution, x = Ae st ,
for the homogenous equation formed for a differential equation. We will not belabor the definition of Laplace transform,
since it can be challenging to evaluate the integral, except
in simple cases. In the future, we will use transform tables,
when needed. However, we can apply our test for linearity,
i.e., doubling the input must double the output, to the transform. The input to the Laplace transform is a function of
time f(t). Doubling the input is expressed as 2f(t).

L {2 f (t )} = 2 f (t ) e

st

dt = 2 f (t ) e dt = 2 F ( s )
st

L dtd f (t ) = sF (s)

L ddt

(7.23)

f (t ) = s n F ( s )

(7.24)

d
transforms to the complex variable s
The operator
dt
which operates on the transformed function of time,
{ f (t )} = F ( s), as a multiplicative factor. The transformation of differential equations to polynomials in the complex variable s allows us to work with algebraic equations,
rather than differential equations.
We will apply the Laplace transform to systems of simultaneous, first-order differential equations, where,

L ddty = L ddtAx

(7.25)

a 1,2 x1

a 2,2 x2

1,1
2,1

(7.27)

s y1 ( s ) a1,1

=
s y2 ( s ) a 2,1

a1, 2 s x1 ( s ) a1,1
a 2, 2 s x2 ( s ) a 2,1

a1, 2
(7.28)
a 2, 2

The convention is to express transformed variables by capital letters, when possible.


The Laplace variable s can be factored out of either form
s y ( s ) a1,1
sY( s ) = AsX( s ) and 1 =
s y2 ( s ) a 2,1

a1,2 s x1 ( s )
a 2,2 s x2 ( s )

to yield

(7.22)

and


(7.26)

yields

a1,1
y (s)
sY( s ) = sAX( s ) and s 1 = s
y 2 ( s )
a 2,1

The resulting transform is doubled.


Our use of the Laplace transform when working with either higher-order differential equations, or with systems of
first-order differential equations, will be to transform differential equations into algebraic equations, neglecting the
initial condition terms. Recall the Laplace transformation
derivatives with respect to time, when the initial conditions
are neglected.

L dtd yy = L dtd aa

sY( s ) = AsX( s )

a1,2 x1 ( s )
a 2,2 x2 ( s )

The Laplace transformation of integration with respect to


1
time introduces the factor, , which is, accordingly, the
s
inverse of the Laplace transformation of differentiation with
respect to time.


L { f (t )dt} = 1s F (s)

(7.29)

7.4 Transpose of a Matrix


The transpose of a vector or a matrix is created by interchanging the rows and columns. The operation is indicated
by a superscript T. For example,


a1
a 2 A T =
a 2

(7.30)

b1
B = B T = b1 b2
b 2

(7.31)

c1,1

C = c1,2
c1,3

c 2,1

c 2,2 (7.32)
c 2,3

A = a1

c1,1 c1,2 c1,3


C=

c 2,1 c 2,2 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.

1
= 1
(7.35)

d 1,1 d 1, 2

D = d 2,1 d 2, 2
d 3,1 d 3, 2

d 1,1 d 2,1
d 1,3

T
d 2,3 D = d 1, 2 d 2, 2

d 3,3
d 1,3 d 2,3

d 3,1
d 3, 2
d 3,3

(7.33)

Note that the elements along the main diagonal of square


matrix C, which have repeated indices, do not move when
the transpose is created, Eq.7.32.

7.5 Matrix Inversion


Consider the following scalar equation of a straight line
y = mx + b
We solve for x by first subtracting b from both sides, and
then dividing both sides by m
y = mx + b

y b = mx

y b
=x
m

This simple bit of scalar algebra is impossible in vectormatrix algebra. Although we can perform the subtraction,
matrix division is not defined!
Although there is no matrix division, there is an analogous operation called matrix inversion. We will develop
the concept with scalar algebra, and then apply it to vectormatrix algebra. Consider the fractional inversion of a scalar
variable. The product of a non-zero scalar and its fractional
inverse is a unity ratio.


= =1

1

for 0

(7.34)

Using the laws of exponents, and recalling that any base


raised to the zeroth power equals 1, a unity ratio can be expressed as the product of a scalar and its fractional inverse.


= 1 = 1 1 = 11 = 0 = 1

(7.36)

The number one is the identity quantity for scalar multiplication. Eq.7.34 is true for any scalar.


1 = 1 =

(7.37)

There is a corresponding identity quantity in vector-matrix


algebra, known as the identity matrix and given the symbol, I, which, when multiplied by a vector or matrix, yields
the original vector or matrix. Although the phrase identity
matrix is used, the identity matrix for a specific vector or,
if matrix is a square matrix, with as many rows and columns
as the largest number of rows or columns in the vector or
matrix, with ones along the main diagonal (top left corner to
bottom right corner) and zeros for every other element.


1
0
I=
0

0
1
0
0

0
0

(7.38)

The simplest case is a square matrix M. A square matrix can


be either pre- or post-multiplied by its identity matrix.

(7.39)
MI = IM = M
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


its corresponding identity matrix is only possible for the
arrangement, which yields the number of rows of the premultiplying quantity equal to the number of columns of the
post-multiplying quantity. For example, row vector A must
be post-multiplied by its identity matrix.
A = a1
AI = a1
= a1

1 0 0 d 1,1 d 1,2

ID = 0 1 0 d 2,1 d 2,2
0 0 1 d 3,1 d 3,2
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


0 d 1,1 + 0 d 2,1 + 1 d 3,1 0 d 1,2 + 0 d 2,2 + 1 d 3,2

a 2
1 0
a 2
= a1 1 + a 2 0 a1 0 + a 2 1
0 1
a 2

d 1,1 d 1,2

D = d 2,1 d 2,2
d 3,1 d 3,2

but
1 0
IA =
a1
0 1

a 2

but

Cannot be evaluated

d 1,1

DI = d 2,1
d 3,1

Whereas, column vector B must be pre-multiplied by its


identity matrix
b1
B=
b 2

The inverse of matrix M, if one exists, is defined as the matrix, which when pre- or post-multiplying M, yields the identity matrix for M. The symbol used for an inverse matrix is
the exponent, 1, e.g., the inverse of M is M1.

1 0 b1 1 b1 + 0 b2 b1
IB =
=
=
0 1 b2 0 b1 + 1 b2 b2

but
b1 1 0
BI =

b2 0 1

Cannot be evaluated

Rectangular matrix C must be post-multiplied by its identity


matrix
1 0
c1,1 c1,2 c1,3
CI =
0 1
c 2,1 c 2,2 c 2,3 0 0

c1, 1 1 + c1, 2 0 + c1, 3 0


=
c 2, 1 1 + c 2, 2 0 + c 2, 3 0
c1, 1 c1, 2 c1, 3
=
=C
c 2, 1 c 2, 2 c 2, 3

d 1,2 1 0 0

d 2,2 0 1 0 Cannot be evaluated


d 3,2 0 0 1

MM 1 = M 1M = I

Matrix inversion is used in vector-matrix algebra to solve


matrix equations. For example, to solve y = Ax for x, we
multiply both sides of the equation by the inverse of A.

0
0
1
c1, 1 0 + c1, 2 1 + c1, 3 0

c1, 1 0 + c1, 2 0 + c1, 3 1


c 2, 1 0 + c 2, 2 1 + c 2, 3 0 c 2, 1 0 + c 2, 2 0 + c 2, 3 0

but
1 0 0
c1,1 c1,2 c1,3
IC = 0 1 0

c 2,1 c 2,2 c 2,3


0 0 1

y = Ax
Cannot be evaluated

Whereas rectangular matrix D must be pre-multiplied by its


identity matrix

(7.40)

A 1y = A 1 Ax

A 1y = Ix

A 1y = x

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

418

Note that an identity matrix I factor can be removed from or


inserted into a vector-matrix equation at will, with the restriction that the resulting matrix product can be evaluated. As a
second example, we will solve the vector-matrix equation,
y = Mx + b for x
y = Mx + b y b = Mx M 1 (y b) = M 1Mx
M 1 (y b) = Ix M 1 (y b) = x

7.5.1 Calculation of an Inverse Matrix


We will preface this discussion of calculating inverse matrices with the reminder, that the inverse of a matrix may not
exist, due to the fact that the system of equations represented
by the matrix equation has no solution. We will also note that
applied mathematicians have put great effort into developing
efficient methods of solving large sets of simultaneous equations. For computational efficiency and the resulting speed,
efficient techniques specifically avoid the task of inverting
a large matrix. Although computer code will avoid inverting matrices, the theoretical equations on which the code is
based must use matrix inversion, since matrix division is not
defined.
Engineering software, including Mathcad and MATLAB,
plus an increasing number of handheld calculators include
functions to calculate the inverse matrices. Although we will
not compute inverse matrices manually, it is important to understand the process of inverting a matrix, because intermediate results in the process are themselves important mathematical entities. The mathematical terminology is somewhat
opaque, in part, because the processes can only be defined
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.
The inverse of matrix A is calculated as the transpose of
the cofactors of A divided by the determinant of A.

A 1 =

a 1,1
a
2,1

a 1,2 a 1,1 a 1,2


a 1,1 a 1,2
=

a 2,2 a 2,1 a 2,2


a 2,1 a 2,2

Fig. 7.1 Graphical procedure for calculating a two by two determinant

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.
The meaning of a determinant is clearest, when we consider the simplest equation, a two by two matrix A, y = Ax
which expands into a set of two simultaneous equations.
y1 a 1,1
y = Ax =
y2 a 2,1

a 1,2 x1
a 2,2 x2

y1 = a 1,1 x1 + a 1,2 x2
y2 = a 2,1 x1 + a 2,2 x2
For a solution x1 and x2 to exist, the two equations must be
independent. Geometrically, in two-dimensional space, the
equations must represent two lines which intersect at one
point. Two cases without any solution occur when the two
lines are identical, and intersect at every point, or are parallel
and have no intersection. The determinant of a set of simultaneous equations equals zero in either of these cases.
Example 1 Two identical equations (or lines)
y1 = 3 x1 + 4 x2
y2 = 3 x1 + 4 x2
a1,1
A =
a2,1

y1 3 4 x1
=

y2 3 4 x2

a1,2 3 4
=
= 3 4 3 4 = 12 12 = 0
a2,2 3 4

Example 2 Two dependent equations (or two parallel lines)

adjoint ( A)
cofactors ( A)T
cof ( A) T
=

determinant ( A) determinant ( A)
A

y1 = 3 x1 + 4 x2
y1 = 3 x1 + 4 x2

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
A =
a 2,1

a1,2 3 4
=
= 3 8 6 4 = 24 24 = 0
a 2,2 6 8

The determinant of a matrix is a scalar. The determinant of


a two by two matrix is easy to remember, and, as the fundamental definition, forms the basis of a recursive algorithm to
calculate the determinant of larger matrices. The determinant
of a two by two matrix is

We have stated without proof that the inverse of the matrix


A is calculated as

a1, 1
A =
a 2, 1

a1, 2
= a1, 1 a 2, 2 a1, 2 a 2, 1
a 2, 2

(7.42)

A 1 =

adjoint ( A)
cofactors ( A)T
cof ( A)T
=

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
determinant

Fig. 7.3 Expansion of the


second minor of a three by three
determinant

Fig. 7.4 Expansion of the


third minor of a three by three
determinant

a 1,1

A = a 2,1
a 3,1

a 1,2
a 2,2
a 3,2

a 1,3

a 2,3 = a 1,1
a 3,3

a 2,2
a
3,2

a 2,3
a 1,2
a 3,3

a 2,1
a
3,1

a 2,3
+ a 1,3
a 3,3

a 2,1
a
3,1

a 2,2
a 3,2

a 1,1

A = a 2,1
a 3,1

a 1,2
a 2,2

a 1,3

a 2,3 = a 1,1
a 3,3

a 2,2
a
3,2

a 2,3
a 1,2
a 3,3

a 2,1
a
3,1

a 2,3
+ a 1,3
a 3,3

a 2,1
a
3,1

a 2,2
a 3,2

a 1,1

A = a 2,1
a 3,1

a 1,2
a 2,2

a 1,3

a 2,3 = a 1,1
a 3,3

a 2,2
a
3,2

a 2,3
a 1,2
a 3,3

a 2,1
a
3,1

a 2,3
+ a 1,3
a 3,3

a 2,1
a
3,1

a 2,2
a 3,2

a 3,2

a 3,2

If there is no solution to the system of simultaneous equations, the determinant of A is zero, and the calculation for
the inverse of matrix A fails, since it is impossible to divide
by zero or, if you prefer, division by zero yields infinity. A
matrix which cannot be inverted is termed singular, a word
used in Victorian England for unique or special and used
in mathematics to mean division-by-zero.
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 minors, repeatedly if necessary, until the determinants are two
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
dropping out both the row and column of the element. The
determinant of a three by three matrix is
a 1,1

A = a 2,1
a 3,1

a 2,2
A = a 1,1
a 3,2

a 2,3
a 1,2
a 3,3

a 1,2
a 2,2
a 3,2
a 2,1
a
3,1

a 1,3

a 2,3
a 3,3
a 2,3
+ a 1,3
a 3,3

a 2,1
a
3,1

a 2,2
a 3,2

The process of expansion by minors is easier to remember


by the patterns, shown in Figs.7.2, 7.3 and 7.4 than by a
formula.
The expansion terms are multiplied by a sign factor calculated by raising minus one to the power equal to the sum of
the row and column of the element, which is the coefficient
of the determinant


+ +
sign factor of
(i + j )
(r + c)
= ( 1)
= ( 1)
= + (7.43)
expansion term
+ +

where i and j are the row and column indices. Note that
the sign factors alternate from positive to negative in a
checkerboard pattern. (Although conventional mathematical notation uses i and j for indices, r and c are useful mnemonics and cannot be confused with the imaginary number
i j 1.)
We will illustrate the calculation of the determinant of a
three by three matrix using two cases. The first case will be
three lines, which form orthogonal axes intersecting at the
origin. Orthogonality, or mutually normal axes, is an important property of engineering systems described by matrix
equations, characterized by a square matrix with non-zero

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

420

element on the main diagonal (top left to bottom right) and


zeros elsewhere.

b1,1 b1,2
b 2,1 b 2,2
B =
b 3,1 b 3,2
b 4,1 b 4,2

y1 2 0 0 x1
y = Px y2 = 0 3 0 x2
y3 0 0 4 x3
This is a diagonal matrix. Diagonal matrices are very important in engineering mathematics because the values of the
elements on the main diagonal are the eigen or characteristic values of the system, described by the set of simultaneous
equations. We are free to expand a matrix by minors, using
a row or column of our choice. Consequently, we always
choose a row or column with as many zeros as possible.
2 0 0
3 0
0 0
0 3
3 0
0
+ 0
= 2
P = 0 3 0 = 2
0 4
0 4
0 0
0 4
0 0 4
P = 2 (3 4 0 0) = 24
The determinant of P is not zero. Therefore, matrix P can be
inverted to solve the equation.
The second example is a system of three simultaneous
equations, which represents two lines that lie in the same
plane; are parallel; and a third line which is normal to the
plane of the two parallel lines. This system of equations has
no solution.

y = Qx

B = b1,1

b 2,2
b 3,2
b 4,2

b 2,3
b 3,3
b 4,3

+ b1,3

b1,3
b 2,3
b 3,3
b 4,3

b1,4
b 2,4
b 3,4
b 4,4

b 2,4
b 2,1 b 2,3
b 3,4 b1,2 b 3,1 b 3,3
b 4,4
b 4,1 b 4,3

b 2,1 b 2,2
b 3,1 b 3,2
b 4,1 b 4,2

b 2,4
b 3,4
b 4,4

b 2,4
b 2,1 b 2,2
b 3,4 b1,4 b 3,1 b 3,2
b 4,1 b 4,2
b 4,4

b 2,3
b 3,3
b 4,3

There is one special case which must be considered. What


is the result, if one takes the determinant of a scalar? This
special case is handled by definition, since the calculation
presented above cannot be performed on a single scalar number. The determinant of a scalar is defined to be the number
itself, i.e.,

determinant ( a ) a = a where a is a scalar (7.44)

7.5.3 Cofactor of a Matrix


The cofactors of a matrix A are defined in terms of the determinant of A. The cofactors are the values of the determinants
of the minors (submatrices), created by deleting the row and
column of each element of the matrix. The cofactor of a matrix A is the matrix of its cofactors. This seemingly circular

y1 2 3 0 x1


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
3
+ 0
= 2
3
+0
Q = 4 6 0 = 2
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
cannot be inverted to solve the equations.
Expanding a matrix larger than three by three is a recursion of this algorithm. A four by four matrix is first expanded
by a row or column of choice into four terms, each including
the determinant of a three by three matrix, which are then
each expanded as illustrated above. Using the first row to
expand the determinant of the matrix B yields

definition requires clarifying illustrations. We will work with


a three by three matrix.
The elements of the cofactor are multiplied by the same
sign factor as the expansion term in a determinant.
sign factor of cofactor element = ( 1) (i + j )
= ( 1)

(r +c)

+ +
= +
+ +

7.5 Matrix Inversion

421

a 1,1

cofactor ( A ) = cofactor a 2,1


a 3,1

a 1,2
a 2,2
a 3,2

a 1,3

a 2,3
a 3,3

a 2,2
( +1)

a 3,2

a 1,2

= ( 1)
a 3,2

+1 a 1,2
( ) a 2,2

7.5.4 Adjoint of a Matrix


The adjoint of the matrix A is the transpose of cofactor( A):
adjoint ( A ) = cofactor ( A )

a 2,2
( +1)

a 3,2

a 1,2

= ( 1)
a 3,2

a
( +1) 1,2

a 2,2

a 2,3
a 3,3

( 1) a 2,1

a 1,3
a 3,3

( +1) a 1,1

a 1,3
a 2,3

( 1) a 1,1

a 2,2 a 2,3
( +1)

a 3,2 a 3,3

a 2,1 a 2,3

= ( 1)

a 3,1 a 3,3

a 2,1 a 2,2

( +1) a
3,1 a 3,2

a 2,2

a 3,2

a 1,2
a 3,2

a 1,2

a 2,2

a 2,3
a 3,3

( +1) a 2,1

a 1,3
a 3,3

( 1) a 1,1

a 1,3
a 2,3

( +1) a 1,1

a 1,3
a 3,3

( +1) a 1,2

a 1,3
a 3,3

( 1) a 1,1

a
a
( 1) a 1,1 a 1,2
3,2
3,1

a
( +1) a 1,1
2,1

3,1

3,1

2 ,1

( 1) a 1,2

3,2

( +1) a 1,1

3,1

3,1

3,1

2,1

2,2

2,1

We will now demonstrate that the expression A 1 =

( 1) a 2,1

a 1,3
a 3,3

( +1) a 1,1

a 1,3
a 2,3

( 1) a 1,1

3,1

3,1

2,1

a 1,3

a 2,3

a 1,3
a 2,3

a 1,2
a 2,2

cof ( A)T
,
A

2 0 0
= 0 3 0
0 0 4

We have calculated the determinant of this matrix and know


that this matrix can be inverted.

a 2,3
a 3,3

( +1) a 2,1

a 1,3
a 3,3

( 1) a 1,1

a 1,3
a 2,3

( +1) a 1,1

3,1

3,1

2,1

a 2,2

a 3,2

a 1,2
a 3,2

a 1,2
a 2,2

2 0 0
3 0
0 0
0 3
0
+ 0
A = 0 3 0 = 2
0 4
0 4
0 0
0 0 4

does indeed yield the inverse of A. To limit the manual computation, we will use a diagonal matrix

[ A]

a 2,3
a 3,3

A = 2

3 0
= 2 (3 4 0 0) = 24
0 4

It remains to calculate cofactor(A).


2 0 0
cof ( A ) = cof 0 3 0
0 0 4

3
( +1)
0

0
cof ( A ) = ( 1)

0
( +1)

0
4

( 1)

0
4

( +1)

0
0

( 1)

3 0
( +1)

0 4

cof ( A ) =
0

(3 4 0 0)

cof ( A ) =
0

0 0

0 4

2 0

0 4
2 0

0 0
0
2 0

0 4

( +1)

0
0
( 2 4 0 0)
0

12 0 0
cof ( A ) = 0 8 0
0 0 6

0 3

0 0

2 0
( 1)
0 0

2 0
( +1)
0 3

( +1)

2 0
( +1)
0 3
0

( 2 3 0 0)

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

422

We now transpose cofactor(A) to yield adjoint(A).


Cofactor(A) is a diagonal matrix and, since the transpose of
a diagonal matrix equals the diagonal matrix, the transpose
of cofactor(A) equals cofactor(A).
T
12 0 0
12 0 0

T
cofactor( A) = 0 8 0 = 0 8 0
0 0 6
0 0 6

The final step is to divide adjoint(A) by determinant(A).

cof ( A )

A =

12 0 0 12
0 8 0

24
0 0 6
=
= 0

24

0
8
24
0

6
24

0
0
0.500

A = 0
0.333
0
0
0
0.250
1

We will now test our result by evaluating the product A 1 A,


which should yield a matrix approximately equal to the identity matrix.

7.6State-Space Representation of Dynamic


Systems
7.6.1 State Variables
One meaning of the word state is condition. A physician
describes the state of a patients health by his or her vital statistics which are variables, such as temperature, blood pressure, and white blood cell count. One could refer to the vital
statistics as state variables, because these are the variables
needed to determine the state of a persons health. Likewise,
the word state also means a condition when used in systems
dynamics. The condition of a dynamic system is defined by
the values of the power variables. For example, the pressure
in a hydraulic system is analogous to blood pressure in a
person.
We have referred to state variables in our study of the
dynamic response of first-order systems. The energy storage variable is the state variable of a first-order system. We
calculated the initial values of power variables in first-order
systems, using the values of the input variable and the energy
storage variable at time, t = 0+ . If we know values of the
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 storage variables are the state variables for the dynamic system.
State variables define the state of the system, because if we

0
0 2 0 0
0.500

0.333
0 0 3 0
A A= 0
0
0.250 0 0 4
0
1

( 0.500 2 + 0 0 + 0 0)

A A = ( 0 2 + 0.333 0 + 0 0)
( 0 2 + 0 0 + 0.250 0)
1

(0.500 0 + 0 3 + 0 0) (0.500 0 + 0 0 + 0 4)
(0 0 + 0.333 3 + 0 0) (0 0 + 0.333 0 + 0 4)
(0 0 + 0 3 + 0.250 0) (0 0 + 0 0 + 0.250 4)

0
0
1

A A = 0 0.999 0 I
0
0
1
1

Success. In practice, it is unlikely that you will ever have


need to manually calculate the inverse of a matrix. Although
it is possible to program the algorithm we just executed, it
is not used in engineering software, because there are much
more computationally efficient methods available, based on
Gaussian elimination.

know these variables as functions of time, we can solve for


any other power variable for any time.
There is an adage that state variables are not unique.
This refers to the ease with which we can use vector-matrix algebra to manipulate a state-space description of a
dynamic system. The fact that we can perform algebra on

7.6 State-Space Representation of Dynamic Systems

v(t)

iR

iL
iC

isource

423
Table 7.1 Power variables in a series RLC circuit

Across Variable

Through Variable

Voltage Source

Known

Resistor

v12

iR

Inductor

v23

iL (state variable)

Capacitor

v3g (state variable)

iC

g
Fig. 7.5 Series RLC circuit showing nodes of distinct voltages between
the elements and positive direction of currents through the elements

the representation of a dynamic system does not change its


underlying properties. We will investigate two very different methods of formulating state equations, and find that
although the state equations produced are very different,
and both describe the same dynamic system, they use different sets of state variables. However, we derive a set of
state equations from the schematic of the energetic model of
a system. The set of state variables will be the energy storage
variables of the system.

7.6.2Example Second-Order Dynamic System


RLC Circuit
There are two power variables in every dynamic element.
The power variables are unknowns, except for one of the two
power variables in a source which we control. Consider the
RLC circuit shown in Fig.7.5. The power variables in the
circuit are shown in Table7.1.

7.6.2.1Second-Order Differential System


Equation of the Series RLC Circuit
We reduced this circuit as an example in Sect.1.5.2.2 to
yield a second-order differential equation relating input v1g
and voltage across capacitor v3g.


d 2 v3 g R dv3 g
1
1
v1g =
+
+
v3 g
LC
L dt
LC
dt 2

(1.48)

If we were interested in another output variable, say the current through the inductor iL, we would need to start from the
equation list and perform a second complete reduction. The
effort of reducing of an equation list to a single differential
equation with the same order as the system increases roughly
as the square of the number of elements in the system. Unfortunately, as reductions become more involved, they also
become more error prone.

There are many advantages of representing a system as


a higher-order differential system equation, introduced in
Chap.2 and to be investigated further in Chap.9. Recall the
algebraic characteristic equation of a system can be written
by inspection of the higher-order differential system equation. The roots of the characteristic equation are the characteristic values, or eigenvalues, of the system, which do
indeed characterize the homogeneous response of the system.
In this chapter, we will develop a technique to derive the
transfer functions of the state variables of a system from its
state-space representation. Since transfer functions can also
be formed from the Laplace transformation of a differential
system equation, we will be able to form the corresponding system equation from a transfer function. We will find
the technique of transforming state equations to transfer
function to system equation to be significantly easier, than
the direct reduction of an equation list to a higher-order differential system equation.

7.6.2.2State-Space Representation of the Series


RLC Circuit
The state-space representation can be thought of as a partial reduction of the equation list to a set of simultaneous
differential equations, rather than to a single higher-order
differential equation. We will use the energy storage variables of a system as its state variable. The state variables of a
system are not unique, because any legitimate linear algebra
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
are important, because of their mathematical convenience in
Modern Control theory, which is based on the state-space
representation.
There are two independent energy stores in RLC circuit:
(1) the capacitor, which stores energy in an electric field, and
(2) the inductor, which stores energy in a magnetic field. The
state variables are the energy storage variables of these two
elements, v3g and iL. The energy storage elements of a system
make the system dynamic. The flow of energy into or out of
a storage element occurs at a finite rate, and is described by
a differential equation, relating the derivative of the energy
storage variable (a state variable) to the other power variable

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

424

v(t)

C
g

Fig. 7.6 Linear graph of the series RLC circuit of Fig.7.5

of the element. The elemental equations of energy dissipaters; transducer and transformer equations; and the continuity
and compatibility equations are all algebraic equations. As
we know from solving for the initial condition of the output
variable in a first-order system, if the values of the state variables are known at any time t, then it is possible to determine
the values of all of the other unknown power variables at
that time.
The state-space representation of a dynamic system separates calculation of the state variables, and calculation of all
of the other power variables into two discrete steps. We will
first formulate the state equations to find state variables v3g
and iL. We will then formulate the output equations to calculate all of the remaining unknown power variables, using
input v, and state variables v3g and iL.
Energetic Equations
Compatibility: v(t ) = v12 + v23 + v3 g
Continuity: i iR = 0
Elemental: v12 = RiR

iR iL = 0 iL iC = 0
dv3 g
diL
iC = C
dt
dt
1 2
1
EC = Cv3 g E L = LiL2
2
2

v23 = L

Energy: E sys = EC + E L

7.6.3 State Equations


State equations are derived by partial reductions of the equation list to a set of first-order differential equations written in
terms of the input variable, the state variables, the elemental
parameters, and time.
The reduction procedure is always the same:
1. Start the reduction for a state equation with the elemental
equation of an energy storage element of the system. You
will have as many state equations, as there are independent energy storages in the system.
2. Rearrange the energy storage elemental equation to place
the derivative of the state variable on the left side by itself
without a coefficient.

3. Proceed to eliminate by substitution all power variables,


except for the input variable and the state variables. The
resulting first-order differential equation is the state equation.
Eliminate the output variables, unless they are also state
variables. We will calculate the output variable after calculating the state variables.
4. Express the state equation in standard form.
The left side:
Only the derivative of the state variable without a coefficient may be on the left side of a state equation.
The state equations must be listed, so that the derivatives
of the state variables are in the same order as the state
variables in the state vector.
The right side:
The coefficients of each state variable are collected, so
that there is only one term for each state variable.
The state variables appear in the same order as the state
vector x.
Constants multiplying state variables and the input variable are distributed.
The input term is last.
List the input variable and the state variables before beginning the reduction. The order of the state variables in the list
will be the order of the state variables in state vector x.
iL
x=
v3 g
Refer the list of state variables, so that you dont lose track
of the variables you wish to keep, as well as those you must
eliminate.
State-Equation Reduction
Input Variable v, State Variables iL , v3 g
First State Equation Begin the reduction with the state
equation for the first state variable, iL. Start with the elemental equation for the inductor.
v23 = L

diL
dt

Rearrange to put the derivative of state variable iL on the left


side by itself without a coefficient.
diL v23
=
dt
L
The variable on the right side, v23, is neither the input nor a
state variable. Eliminate it by substitution.
diL 1
= v v12 v3 g
dt
L

7.6 State-Space Representation of Dynamic Systems

425

Eliminate v12 by substitution because it is neither the input


nor a state variable.
diL 1
= v RiR v3 g
dt
L

Eliminate iR by substitution because it is neither the input nor


a state variable.
diL 1
= v RiL v3 g
dt
L

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
that the state variables appear in the state vector, followed by
the term with the input variable.
di
diL 1
1
R
1
1
R
= v iL v3 g L = iL v3 g + v
dt
L
L
L
dt
L
L
L
Second State Equation Begin with the elemental equation
of the capacitor.
iC = C

dv3 g
dt

Rearrange to put the derivative of state variable v3g on the


left side.
dv3 g

i
= C
dt
C

Eliminate iC since it is neither the input nor a state variable.


dv3 g
dt

iL
C

If we know input v and the state variables, iL and v3g, over a


period of time, then we can calculate every unknown power
variable in the system over the same period. Output equations are formulated in terms of the input and the state variables. The calculation sequence is fixed by necessity. The
state equations must be solved before the output equations.
Solution of the state equations requires a technique for solving a set of simultaneous differential equations. The solution
of the output equations is purely algebraic. Consequently,
only use the algebraic equations of the systems energetic
equations. Never use the elemental equations for the energy
storage elements, since these are differential equations.
Example Output Equation Derive the output equation for
the voltage across inductor, v23. Identify the input, output,
and state variables, before beginning a reduction.
Input v, Output v23, State Variables iL, v3g
We will start the reduction with a compatibility equation.
v = v12 + v
23 + v3g

Summary of the State Equations List the state equations


so that the derivatives of the state variables are in the same
order as the state variables in state vector x. This is the form
we will use for the vector-matrix notation.
diL
1
1
R
= iL v3 g + v
dt
L
L
L
dv3 g 1
= iL
dt
C

7.6.4 Output Equations

Input
Known

This is a state equation in standard form with v3 g and iL as


state variables.

variables. The two equations are coupled to one another.


Both must be satisfied, if either is to be true. We will use the
RungeKutta numerical method to solve sets of simultaneous differential state equations. The RungeKutta method is
available in Mathcad and MATLAB as built-in functions. We
will also program the algorithm in MATLAB.
The state equations must be solved before the output
equations are solved. Solution of state equations requires
a numerical method for solving simultaneous differential
equations, such as the RungeKutta method. Solution of the
output equations will be purely algebraic, in terms of the
state variables and the input variable.

The state equations are a set of simultaneous first-order differential equations, because both equations share the state

State
Variable
Known

Rearrange to express the desired unknown on the left. Place


the power variables to be eliminated, the state variables (in
the order of the state vector), and input on the right.
v
= v12 v3g + v
23

Input

Desired
Unknown

(7.45)

Desired
Unknown

State
Variable
Known

Known

The voltage drop across resistor v12 is neither the desired


output nor a state variable. Eliminate v12 by substitution.
There is only one equation which contains v12, other than the
compatibility equation. That is the elemental equation for the
resistor

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

426

v
= RiR v3 g + v
23

Input

Desired
Unknown

State
Variable
Known

Known

The current through the resistor is neither the desired output


nor a state variable. Eliminate it by substitution. Use the continuity equation for node 2
v
= R iL v3 g + v
23

Input

Desired
Unknown

State
Variable
Known

State
Variable
Known

iC = iL
iR = iL
i = iL
The output equation for the voltage drop across the resistor
starts with the elemental equation for the resistor. One substitution is needed to eliminate iR.
v12 = RiR v12 = RiL
Summary of the output equations for the unknown power
variables
iR = iL
i = iL
v12 = RiL
v23 = RiL v3 g + v

The general vector-matrix form of the state equations is




dx
= Ax + Bu
dt

(7.46)

x is the vector of state variables, a.k.a. the state vector,


iL
x=
v3 g

Known

We have eliminated every variable except the input, the state


variables, and the desired output variable. Consequently, this
is the output equation.
The standard form for output equations is similar to that
of the state equations, with one crucial difference. Output
equations must be purely algebraic. There cannot be a derivative in an output equation! The state variables must be
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
on the right side.
Make sure that state variables are in the same order in
the output equations, as they are in the state equations. If the
state variables are not in the same order, then the two vectors
are not the same state vector. The order of the output equations themselves is arbitrary, unless an output vector y was
specified.
We will now formulate the output equations for the remaining unknown power variables in the system, iC , iR , the
current i from the source, and v12.
Input Variable v, State Variables iL , v3 g
The output equations for iC , iR , and i are trivial, since all
of the currents are equal to the current through the inductor,
which is a state variable.

iC = iL

7.6.5 Vector-Matrix Form of the State Equations

Matrix A consists of expressions of the elemental parameters, known as the coefficient matrix, because its elements
are the coefficients of the state variables in the state equations. Vector (or matrix) B is the input matrix, since it is
multiplied by the vector of inputs, u. In this example, we
have just one input, so u is a single function, the input v(t ),
not a vector of functions.
u = v (t )
State equations for the RLC system are
diL
1
1
R
= iL v3 g + v
dt
L
L
L
dv3 g 1
= iL
dt
C
which are written in the matrix form by reversing the process of vector-matrix multiplication. Coefficient matrix A is
always square, with as many rows and columns, as there are
state variables.
Check the order of state equations, and the order of terms
in the state equations are the same as the order of state variables in the state vector, before transcribing the state equations into vector-matrix form. It reduces errors to write state
vector x on both sides of the equation first
dx
= Ax + Bu
dt

d iL

dt v3 g

iL
+ v
v
3 g

before you enter the elements of coefficient matrix A and


input matrix B. Missing terms correspond to elements with a
coefficient of zero.
dx
= Ax + Bu
dt
R
L
=
1
C

d iL

dt v3 g

1
i
1
L L
+ L v

v
0 3g 0

7.6 State-Space Representation of Dynamic Systems

427

7.6.6Vector-Matrix Form of the Output


Equations

dx
= Ax + Bu
dt
y = Cx + Du

The general matrix form of the output equations is


(7.47)
y = Cx + Du
where y is the vector of output variables, x is the vector of
state variables, and u is the vector of inputs. Matrix C is
the output matrix, which has as many rows as there are output variables (or output equations), and as many columns as
there are state variables. Matrix D has the peculiar name of
direct pass-through matrix. The effect of D is to scale the
value of the input and add the scaled input value to the output
variables without passing through the state equations.
The output equations for the RLC circuit are
iC = iL
iR = iL
i = iL
v12 = RiL
v23 = RiL v3 g + v
Again, write state vector x and output vector y first.

y = Cx + Du

iC
i
R
i =

v12
v23

iL
v +
3g

Then, enter the elements of output matrix C and the direct


pass-through matrix D to reduce error.

y = Cx + Du

0
iC 1
0
i 1

0
0
R
iL
i = 1
0 + 0 v

v3 g

0 0
v12 R
v23 R 1
1

Matrix C is the output matrix. Matrix D is the passthrough matrix, so named because the product Du contributes to the output vector, independent of the state equations.

7.6.7Numerical Solution of the State


and Output Equations
Note that state equations and output equations are different
types of vector-matrix equations.

State Equations
Output Equations

State equations are simultaneous first-order differential


equations. Output equations are simultaneous algebraic
equations, not differential equations.
What can be lost in the convenience of vector-matrix
shorthand notation is the shape of each of the matrices,
A, B, C, and D. There are as many state equations as there
are state variables. Consequently, coefficient matrix A is a
square matrix with as many rows as columns. It is possible
to formulate state equations to calculate the response of systems without an input. Such a system would respond to energy in the system at time, t = 0+ . Alternatively, it is possible
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,
a matrix filled with elements equal to zero, or simply absent.
If there are as many inputs as state variables, then matrix B
is square. If there are more inputs than state variables, then
there are at least two input matrices, B1 and B2.
Similarly, there may be only one output equation, or more
output equations than there are state variables. Consequently,
output matrix C may be a row matrix or a rectangular matrix.
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
system at any moment in time. State equations are first-order
differential equations, since state vector x is differentiated
once with respect to time. In the general case, a state equation relates the rate of change of a state variable, xi, to the
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
equations must be solved simultaneously, to determine the
future state of the system.
We will solve state equations for the values of state variable vector x at discrete instances in time, using a numerical
algorithm, the RungeKutta method. The syntax of engineering mathematical software, e.g., Mathcad and MATLAB, requires that state and output equations be expressed in vectormatrix form. The numerical solution of the state equations
produces the state variables history over the duration of the
solution. The result of the RungeKutta method is presented
as an array, where the first column contains the instances in
time at which the solution was calculated, and the remaining
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
time. The temporal sequences of the state variables, that
is, their values over time, are then used to evaluate the output
variables. If we were analyzing a mechanical system, and we
were interested in either displacement or acceleration (not
the state variable, velocity), we can easily integrate or differentiate the temporal sequences numerically. Numerical inte-

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

428
Fig. 7.7 Schematic of a thirdorder rotational system driven by
a velocity source

Bearing Damping b 2
Pinion
N1 teeth

Inertia J2

Compliant Shaft
Spring Constant K
Inertia J1
Fluid Coupling
Damping b1

Gear
N 2 teeth

(t)
Angular velocity source

gration is both easy to program in MATLAB and accurate.


Differentiation, although easy to program, is less accurate, as
we shall see in the next chapter.

Table 7.2 Parameter values


Damping b1

b1 =

30 N m sec
Spring Constant K
rad

K =5

Damping b2

b2 =

4 N m sec
rad

Number of Pinion
Teeth, N1

N1 = 16

Inertia J1

J1 = 1kg m 2

Number of Gear
Teeth, N2

N 2 = 32

7.7.1Third-Order Dynamic System Example:


A Rotational Mechanical System

Inertia J2

J 2 = 6 kg m 2

A rotational mechanical system is shown schematically in


Fig.7.7. The system consists of an angular velocity source
acting on a fluid clutch with damping b1. The clutch drives
a rotational inertia J1, which is attached to a compliant shaft
with spring constant K. The shaft drives the pinion of a gear
set. The output shaft of the gear drives inertia J2, and is supported on bearings with damping b2. Derive the state equations and output equations for this system, where the output
variables are acting to accelerate rotational inertias J1 and J2,
using the parameter values in Table7.2.
First, find the nodes of distinct values of the across variable, angular velocity, and draw the linear graph.
Derive the transducer equations for the pinion and gear.
Use the geometry of the motion to derive the first transformer equation, then sum power flows into the transformer to
derive the second transformer equation. The pinion and gear
must have the same translational velocity where they mesh.
Use the tooth counts as a measure of the circumferences. Remember, the pinion and gear counter-rotate.

The power flows through the two branches of the transformer interface sum to zero.

7.7Example Derivations of State and Output


Equations

Ppinion + PGear = 0 TP 3 g + TG 4 g = 0

N1
3 g = 0
N2

3 g + TG

rpinion N1 and rgear N 2


N1 3 g = N 2 4 g

N
4 g = 1 3 g
N2

N1
TG
N2

Energetic Equations
Continuity:

T = Tb1 Tb1 = TJ1 + TK TK = TP TG TJ 2 Tb2 = 0
Compatibility:

1g = 12 + 2 g 2 g = 23 + 3 g 4 g = 4 g
Elements:
Tb1 = b1 12

Tb 2 = b 2 4 g

TJ1 = J1

rpinion 3 g = vmesh = rgear 4 g


where

TP =

d 2g

TJ 2 = J 2

dt

dTK
= K 23
dt
d 4g
dt

Energy:

E sys = E J + E K + E J
1

1
E J1 = J1 22 g
2

EJ

1
= J 2 22 g
2

EK =

TK2
2K

Nm
rad

7.7 Example Derivations of State and Output Equations

429

Fig. 7.8 Schematic of a thirdorder rotational system showing


nodes of distinct values of the
across variable rotational velocity

Bearing Damping b 2
Pinion
N1 teeth

Inertia J2

Compliant Shaft
Spring Constant K

Inertia J1
Fluid Coupling
Damping b1

(t)

Gear
N 2 teeth

Angular velocity source

Reduce the equation list to state and output equations. Define the state variable before beginning the reduction to limit
errors:

Second State Equation


dTK
= K 23
dt

2 g

x = TK
4 g

d 2g
dt

1
Tb TK
J1 1

dt
dt

b1
J1

dt

d 2g
d 2g

TJ 2 = J 2

2g

( (

d 2g
dt
d 2g
dt

1
TJ
J1 1

1
b1 12 TK
J1

1
b1 2 g TK
J1

d 4g

State Equation 2

d 4g

dt

1
TG Tb2
J2

d 4g

Third State Equation

dt

First State Equation

dt

dTK
N
= K 2g + K 2 4g
dt
N1

d 4g

State Equations:
Input 1g State variables 2 g , T K , 4 g

d 2g

dTK
= K 2 g 3 g
dt

dTK
N
= K 2 g + 2 4 g
N1
dt

Derive the state equations by starting with the elemental


equation with the derivative of corresponding state variable.
Move the derivative to the left side, and clear its coefficient.
Proceed to eliminate all variables from the right side, which
are neither state variables nor the input variable. The right
side must be purely algebraic. No derivatives appear on the
right side.

TJ1 = J1

dt
d 4g

b2
1
TK 4 g
J2
J2

dt

1
TJ
J2 2

1
TK b 2 4 g
J2

State Equation 3

Summary of State Equations

b1
1
TK + State Equation 1
J1
J1

d 2g
dt

b1
J1

2g

b1
1
TK +
J1
J1

dTK
N
= K 2 g + K 2 3g
dt
N1
d 4g
dt

b2
1
TK 4 g
J2
J2

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

430
Fig. 7.9 Linear graph of the
third-order rotational system of
Fig.7.8

b1

2 g
d

TK =

dt

4 g

1
J1
0

1
J2

0
b1
2g

J
N2

TK + 1
K

N1
0

4
g
0
b2

J2

TJ 2 =

TJ1 = b112 TK

TJ1 = b1 2 g TK

TJ1 = b1 2 g TK + b1

N2
TP b 2 4 g
N1
Output Equation 2

Summary of output equations


TJ1 = b12 g TK + b1
TJ 2 =

N2
TK b2 4 g
N1

Write the output equations in vector-matrix form. First, write


output vector y and state vector x.

y = Cx + Du

b
TJ1 1
=
TJ 2 0

TJ1 = Tb1 TK

TJ 2 = TG Tb2

TJ1
=
TJ 2


2g
TK +

4g

Then enter the components of output matrix C and the passthrough matrix D.

Output Equation for TJ1


Tb1 = TJ1 + TK

N2
TK b2 4 g
N1

Output Equations:
Input 1g , State Variables 2 g , TK , 4 g,
Output Variables TJ1, TJ 2.

Output Equation for TJ 2

Then write the elements of coefficient matrix A and input


matrix B.
b1

J
2 g 1

TK = K

dt

4 g
0

TG J2 b2

TJ 2 =

2 g


TK +

4 g

TP

TG TJ 2 Tb 2 = 0

dx
= Ax + Bu
dt

J1
g

Write the state equations in vector-matrix form. Write the


state vectors first.

(t)

-N1
4g= ___ 3g
N2

Output Equation 1

1
N
2
N1

b1 2 g
T + b1 t
()
b 2 K 0 1g
4 g

7.7.2Fourth-Order Dynamic System Example:


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 masses 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, Table7.3. From

7.7 Example Derivations of State and Output Equations

x,v

K1
M1

b
M1

K2

x,v

K1

M2

F(t)

431

b
g

M2

F(t)

K2

Fig. 7.10 Schematic of a fourth-order translational system driven by


a force source

Fig. 7.11 Schematic of the translational system, annotated with nodes


of distinct values of the across variable, translational velocity

the Newtonian perspective, this mechanical system is classified as having two degrees of freedom, since each mass can
displace independently in the x direction. Note that the positive x direction is indicated on mass M2 to correspond to the
direction of input force F(t).
The first step of our analysis is independent of the size or
type of the system; whether we will represent the system as
a single higher-order differential equation; or in state-space
as a set of simultaneous first-order differential equations. We
must find the nodes of distinct values of the across variable,
translational velocity, Fig.7.11, so that we can draw the linear graph, Fig. 7.12.

2. The state variable terms appear in the order of the


state vector,
3. Constants multiplying state variables are distributed,
4. The input term is last.
We will decide the order of the state variables before beginning the reduction, in order to express each state equation in
the form we can transcribe directly into vector-matrix notation. The order of the state variables in the state vector is
arbitrary but once defined must be respected. We will define
the state vector as:
v1g
v
2g
x=
FK1

FK2

State Vector:

Energetic Equations
Continuity: FM1 FK1 = 0 F + FK1 FM 2 FK2 Fb = 0

Input F, State Variables v1g , v2 g , FK1, FK2

Compatibility: v1g = v12 + v2 g v2 g = v2 g


Element:
dFK1
dt

FM1 = M 1
= K1v12

dv1g
dt
dFK2
dt

FM 2 = M 2
= K 2 v2 g

dv2 g

(1) FM1 = M 1

dt

dv1g

Fb = b v2 g

dt

dv1g

EM =
1

FK22

1
1
M 1v12g E M 2 = M 2 v22g E K1 =
E K2 =
2
2
2 K1
2K2

(2) FM 2 = M 2
dv2 g
dt

Reduction to State Equations


Reminders:
The derivation of a state equation begins with the energy
storage element equation with the derivative of that state
variable.
Only the derivative of the state variable without a coefficient may be on the left side of a state equation.
Eliminate by substitution all power variables on the right
side, except the state variables and the input variable.
Once elimination by substitution is complete, then write
the state equation in standard form on the right side where:
1. The state variables are collected,

dv2 g
dt

dt

1
FK
M1 1

Energy: E Sys = E M1 + E M 2 + E K1 + E K2
FK21

dt

dv1g

dv2 g
dt

1
FM
M1 1

State Equation1
dv2 g

dt

1
( F + FK1 FK2 Fb )
M2

1
( F + FK1 FK2 bv2 g )
M2

1
1
1
b
v2 g +
FK
FK +
F
M2
M2 1 M2 2 M2
State Equation 2

(3)

dFK1
dt
dFK1
dt

(4)

dFK2
dt

= K1v12

dFK1
dt

= K1v1g K1v2 g
= K 2 v2 g

= K1 (v1g v2 g )
State Equation 3

State Equation 4

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

432

Table 7.3 Power variables in the fourth-order mechanical system

K1

F(t) b

M1
g

K2

M2

Across variable

Through variable

Force Source

v 1g (state variable)

Known

Mass M1

v 1g (state variable)

FM1

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)

Damper b

v 2 g (state variable)

Fb

Fig. 7.12 Linear graph of the fourth-order translational system

Summary of the State Equations


dv1g
dt
dv2 g
dt
dFK1
dt
dFK2
dt

1
FK
M1 1

1
1
1
b
v2 g +
FK1
FK2 +
F
M2
M2
M2
M2

= K1v1g K1v2 g
= K 2 v2 g

Express the state equations in matrix form


dx
= Ax + Bu
dt
v1g
v
d 2g
=
dt FK1

FK2

v
1g


d v2 g
=0
dt FK1

FK2 K1
0

v1g

v2 g
F +
K1
F
K2
0

1
M1

M2

1
M2

K1
K2

0
0

0
0
v
1g 1

1 v2 g

+ M2 F

M 2 FK1


0
0 FK2
0

Output Equations
The state equations are solved, yielding the values of state
variables over the period of interest, before the output variables are solved. The state variables and input variable are
then used to formulate the output variables equations. In this
system, after the state variables are known, the only remaining unknown power variables are the forces acting to acceler-

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 .
The output variable equations are always purely algebraic.
There can be no time derivatives in the output equations.
We will use the state variable vector and input variable to
calculate the output variables. Therefore, the standard form
of the output equation places the output variable on the left
side without a coefficient, and these conditions apply to the
right side.
The state variables are collected,
The state variable terms appear in the order of the state
vector,
Constants multiplying state variables are distributed, and
The input term is last.
Derive the output equations for FM1, FM 2, v12, Fb
(1) Output Equation for FM1
FM1 FK1 = 0 FM1 = FK1
(2) Output Equation for FM 2

F + FK1 FM 2 FK2 Fb = 0
FM 2 = F + FK1 FK2 Fb
FM 2 = F + FK1 FK2 bv2 g
FM 2 = bv2 g + FK1 FK2 + F
(3) Output Equation for v12
v1g = v12 + v2 g

v12 = v1g v2 g

(4) Output Equation for Fb

Fb = b v2 g

Summary of the output equations for FM1, FM 2, v12, Fb


FM1 = FK1
FM 2 = bv2 g + FK1 FK2 + F
v12 = v1g v2 g
Fb = bv2 g

7.7 Example Derivations of State and Output Equations

433

Fig. 7.13 Schematic of a hybrid


rotational mechanical-fluidtranslational mechanical system
driven by a torque source

C
T(t)

R2

Pump

FM1


FM 2

y = Cx + Du
=
v12
Fb

v1g


v2 g

+
FK1
F
K2

FM1 0 0 1 0 v1g 0


FM 2 0 b 1 1 v2 g 1

=
+ F
y = Cx + Du
v 1 1 0 0 FK1 0
12

Fb 0 b 0 0 FK2 0

We relate volume Vol displaced per revolution to the volume


flow rate from the pump QP, and the angular velocity of the
shaft driving the pump 1g .
QP =

A fourth-order fluid-mechanical system of Sect.6.4.4.3 is


shown schematically in Fig.7.13. This is a hybrid system
with three different energy domains; rotational mechanical, fluid, and translational mechanical. The gear pump and
the hydraulic piston are transducers which interface the dissimilar energy domains.
The only difference in the process of expressing a system
with transducers or transformers in state-space from that of
a system with a single energy domain is that the transducer
or transformer equations must be derived and included in the
equation list. The first step, as always, is to find the nodes of
distinct values of the across variables.
Draw the linear graph by first drawing the interfaces for
the two transducers, and then adding the remaining elements
between their respective nodes. The two branches of an interface represent only the transformation or transduction of
power. Do not assign them properties of energy storage or
dissipation.
Recall we derive the transducer equations for a shaft-driven positive displacement pumps, using the fact that volume
Vol of fluid was displaced with each revolution of the shaft.

Vol
1g
2

The second transducer equation for the gear pump was derived by setting the sum of the two power flows into the gear
pump as equal to zero. These power flows are the products of
the through and across variables of the two branches, which
form the interface of the linear graph transducer symbol.

Ppump + Ppump = 0 TP 1g + QP p2 g = 0
mech

7.7.3Fourth-Order Dynamic System Example:


A Fluid-Mechanical System

R1

Express the output equations in matrix form.

Hydraulic
Piston
Area A

fluid

We then eliminated the volume flow rate of the pump QP by


substitution
TP 1g +

Vol
1 g p2 g = 0
2

TP 1g +

Vol
1g p2 g = 0
2

to yield the second transducer equation of the pump


TP =

Vol
p2 g
2

The transducer equation for the hydraulic piston cylinder


was derived by calculating the net pressure force acting on
the fluid-side face of the piston, and equating that to the
force acting on the mechanical-side face.
FA = Ap4 g
The second transducer equation for the hydraulic piston cylinder was derived, by summing the power flows into the hydraulic piston

Ppiston + Ppiston = 0 QA p4 g + FA v5 g = 0
fluid

mech

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

434
Fig. 7.14 Schematic of a hybrid
rotational mechanical-fluidtranslational mechanical system,
annotated to show distinct
locations of the across variables:
rotational velocity, pressure, and
translational velocity

and then substituting in the relationship for FA:


QA p4 g + A p4 g v5 g = 0

QA p4 g + A p4 g v5 g = 0

T(t)
b

Pump

QR 2 QI = 0

QP QR 1 QR 2 = 0

QI Q C QA = 0

Compatibility: 1g = 1g
p4 g = p4 g
Vol
1g
Element: QP =
2
FA = Ap4 g

p2 g = p23 + p34 + p4 g

1g
Q
I
x=
p4 g

v5 g

p23 = R 2 QR 2

QC = C

dp4 g
dt

Reduction to State Equations


Input T State Variables 1g , QI , p4 g , and v5 g
(1)

TJ = J

d 1g

FM = M

EI =

d 1g

d 1g

dt

dt

p34 = I

dQI
dt

d 1g
dt

dv5 g

d 1g

dt

1 2
1
IQI EC = C p42g
2
2

dt

EM =

1
Mv52g
2

d 1g
dt

1
TJ
J

1 Vol
1

R1QR 1 + T
J 2
J

1 Vol
1

R1 QP QR 2 + T
J 2
J

1 Vol
1 Vol
1

R1QP
R1QR 2 + T
J 2
J 2
J

1 Vol
1
1 Vol Vol
1g
R1QI + T

R1
J 2
J 2
J
2

d 1g
dt

d 1g 1
1
1 Vol
T TP )
= T+
(
p2 g
J
dt
J
J 2

dt

dt
1
J 12g
2

d 1g

d 1g

Energy: E sys = E J + E I + EC + E M

EJ =

Vol
TP =
p2 g
2
TJ = J

Identify state variables of the system. The state variables are


the energy storage variables. Define the order of state variables in the state vector.

dt

v5 g = v5 g

QA = Av5 g

p2 g = R1QR1

FA FM = 0

R1

The linear graph sign convention should be respected during


the derivation of the state equations. The signs can then be
fixed, so that a positive input yields a positive output.

Continuity: T TJ TP = 0

2 R2 3 I

QA = Av5 g .

Energetic Equations

Hydraulic
Piston
Area A

R1 Vol 2
R1 Vol
1

1g

QI + T
J 2
J 2
J
State Equation 1

7.7 Example Derivations of State and Output Equations

435

Fig. 7.15 Linear graph of the


hybrid rotational mechanicalfluid-translational mechanical
system shown in Fig.7.14

Vol
Q P = ___ 1g
2
1

T(t) J

R2

TP

FA= Ap4g

I
3

QP R1

QA

p34 = I

(2)

dQI
dt

dQI 1
= p34
dt
I

dQI 1
=
p2 g p23 p4 g
dt
I

(4) FM = M

dt

dp4 g

dp4 g
dt

dp4 g
dt

dt

dp4 g
dt

A
1
QI + v5 g
C
C

1
FM
M

A
p4 g
M

dv5 g
dt

1
FA
M

State Equation 4

R1 Vol 2
R1 Vol
1

1g
QI + T
J 2
J 2
J

Express the state equations in matrix form,

1
QC
C

dp4 g 1
1
QI QA )
QI + Av5 g
=
(
C
dt
C
=

dt

1
A
QI + v5 g
C
C
A
=
p4 g
dt
M

State Equation 2

dv5 g

dt
dv5 g

R1 Vol
R1 + R 2
dQI
1
QI p4 g
1g
=

dt
I 2
I
I

dp4 g

R1 Vol
R1 + R 2
dQI
1
1g
QI p4 g
=

dt
I 2
I
I

R1 Vol
dQI
1
R2
1g QI
QI p4 g
=

dt
I
I
I
2

QC = C

dt
dv5 g

d 1g

R1
R2
dQI
1
QI p4 g
=
QP QR 2
dt
I
I
I

(3)

Summary of the State Equations

R1
R2
dQI
1
=
QR 1
QR 2 p4 g
dt
I
I
I

dv5 g

dt

FA
g

dQI 1
= R1QR 1 R 2 QR 2 p4 g
dt
I

State Equation 3

R1 Vol 2

J 2

1g R Vol
1

d QI I 2
=
dt p4 g


0
v5 g

R1 Vol

J 2

R + R2
1
I

1
C
0

1
I

A
M

1
1g
0 Q J
I + 0 T

A p4 g 0
v

C 5 g 0

dx
= Ax + Bu
dt

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

436
Table 7.4 Power variables in
the fourth-order fluid-mechanical
System

Across variable
Torque source

1g (state variable)

T (input)

Inertia J

1g (state variable)

TJ

Mechanical side of gear pump

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

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

Output Equations
The power variables of this system are shown in Table7.4.
Remember, the output equations are algebraic equations, in
terms of the input and state variables. Do not include derivatives on either side of any output equation.
(1) Output Equation for TJ :
T TJ TP = 0
TJ =

Vol
p2 g + T
2

Vol
TJ =
R Q +T
2 1 R1

TJ =

Vol
R1QR1 + T
2

Vol
TJ =
R ( QP QR2 ) + T
2 1

Vol Vol
Vol

TJ =
R
R Q +T
2 1 2 1g 2 1 R2
2

Vol
Vol
TJ =
R
R Q +T
2 1 1g 2 1 R 2
(2) Output Equation for TP:
T TJ TP = 0 TP = TJ + T
Vol

Vol
TP =
R1 1g
R1QR2 + T + T

2
2

Vol
p2 g =
R R1QR 2
2 1 1g
(4)Output Equation for QR1:
QR1 =

p2 g
R1

QR1 =

1 Vol

R1 1g R1QI

2
R1

Vol
QR1 =
QI
2 1g
(5)Output Equation for QR2: QR2 = QI
(6)Output Equation for p23: p23 = R 2 QR2 p23 = R 2 QI
(7)Output Equation for p34:
p2 g = p23 + p34 + p4 g p34 = p2 g p23 p4 g
V
p34 = R1 1g ( R1 + R2 ) QI p4 g
2
(8)Output Equation for QC:
QI QC QA = 0 QC = QI QA

Vol
Vol
TP =
R +
RQ
2 1 1g 2 1 R2

(state variable)

(3)Output Equation for p2 g:


Vol
2
TP =
p2 g p2 g = TP
2
Vol
Vol 2
2
Vol
p2 g =
R 1 1 g +
R1QR 2

2
2
Vol

TJ = TP + T

Vol
Vol
TJ =
RQ
R Q +T
2 1 P 2 1 R 2

Through variable

QC = QI + Av5 g
(9) Output Equation for FA: FA = Ap4 g
(10) Output Equation for FM:
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
position axes.
A two-dimensional geometric space, i.e., a plane, is
formed by defining two coordinate axes, a three-dimensional
space by defining three coordinate axes, a four-dimensional
space by defining four coordinate axes, etc. Although we
cannot visualize spaces above three dimensions and call such
spaces hyperspace, which means beyond space, hyperspace follows the same mathematical rules as does two- and
three-dimensional spaces. You can define a space to have as
many dimensions as you want. The only restriction is that
its axes must be independent. Axes are independent, if you
cannot plot a point on one axis using the other axes, except at
the origin. In the most fundamental sense, a space is defined
by its coordinate coordinates. If we define coordinates, then
we define a space. The axes of the space do not need to have
the same units. We often use two-dimensional spaces with
time or position as the horizontal axis and a different physical quantity as the vertical axis.
A state-space is a space defined by coordinates, which
are the state variables of a system. It is neither a real

Vol
Vol
TJ =
R
R Q +T
2 1 1g 2 1 I
2

Vol
Vol
TP =
R +
RQ
2 1 1g 2 1 I
Vol
p2 g =
R R1QI
2 1 1g
Vol
QR1 =
QI
2 1g
QR2 = QI
p23 = R 2 QI
Vol
p34 =
R ( R1 + R2 ) QI p4 g
2 1 1g
QC = QI + Av5 g
FA = A p4 g
FM = A p4 g
Output equations in matrix form, y = Cx + Dy
2

Vol
R

2 1

Vol
TJ

R1
T
2
P
Vol
p2 g

2 R1
QR1


Vol
QR2 = 2
p
23
0
p34
0
Q
C Vol
FA
R1

2
FM
0

Vol
R

Vol

R1
2

R1

1
R2

(R

+ R2

1
0
0

7.8 Why State-Space is called State-Space


State-Space is a very impressive term for a simple concept.
Mathematically, a space is defined by the coordinate axes,
which allow you to plot a vector in that space. All planes
and spaces you have worked with in engineering are abstractions, with the exception of purely geometric spaces used
to draw parts. You used force spaces in statics to perform
equilibrium analysis. The coordinates of a force space are the
force components, Fx, Fy, and Fz. Although an abstraction,
it is easy to visualize a force vector in a three-dimensional

0
0

1
0
A
A

1
0


0
0


1g 0

0 QI 0

+ T
p3 g 0

0
v4 g 0


0
0

0
A

0
0
0

space in a geometric sense, but nor is a force space. The energy storage variables satisfy the mathematical requirement
for axes, because they are independent. The values of the
state variables at any moment in time are the coordinates of
a point (a vector from the origin) in the state-space of that
dynamic system. We can, and do, define hyper state-spaces.
A dynamic system with four independent energy storage
elements has four state variables. Consequently, its statespace is four dimensional, because its state vector has four
components. Is it easier to work with three-dimensional
state-spaces than four-dimensional state-spaces? Not re-

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

438

ally. It makes little difference how many dimensions there


are in the state-space in practice, because our visualization
will be two-dimensional plots of individual state variables
or output variables vs. time. Our major concern with higher-order systems will be with the accuracy of our model,
which will degrade with increased complexity of higherorder systems.

7.9Expression of Systems Equations


in State-Space
The set of energy storage variables will be our preferred
choice as the state variable of a system, when we derive the
state equations starting with a schematic or linear graph. Energy storage variables have physical meaning. The derivation
of state equations from the equation list is straightforward.
However, other sets of variables can be chosen as the state
variables of a system.
We will now develop an algorithm to express higher-order system equations and transfer functions as sets of state
equations. The technique is straightforward if the input is not
differentiated. Systems with derivatives of input variables
require a more complicated algorithm.

State variables are defined recursively as the output variable


and its derivatives on the right side.
x1 y

dy
dt
d2y
x3 2
dt
d3y
x4 3
dt
x2

(7.51)

All but one state equation is formed from the definitions of


the state variables, by noting the relationships between the
state variables and the derivatives of the output variable.


x1 y
dy dx1
=
dt
dt
2
d y dx
x3 2 = 2
dt
dt
3
d y dx
x4 3 = 3
dt
dt

x2

(7.52)

The final state equation is the equation

7.9.1Algorithm to Express a Higher-Order


System Equation Without Differentiation
of the Input as State Equations

b
a d 3 y a d 2 y a dy a4
d 4 y
y+ 4u
= 1 3 2 2 3

4
dt
a0
a0 dt
a0 dt
a0 dt a0

If the input is not differentiated, then a set of state equations is defined as the output variable and its n-1 derivatives,
where n is the order of the differential equation. We will use
a fourth-order system equation for illustration, where the
input variable is u (t ), and the output variable is y (t ).

noting that

d4y
d3y
d2y
dy
a0 4 + a1 3 + a2 2 + a3
+ a4 y = b4 u (7.48)
dt
dt
dt
dt

Standard form for higher-order system equations is to clear


the highest-order derivative of the output variable of any coefficient term.
 d 4 y a1 d 3 y a2 d 2 y a3 dy a4
b
+ 3 + 2 +
+ y = 4u
4
dt
a0 dt
a0 dt
a0
a0 dt a0
(7.49)
Isolate (solve for) the highest-order derivative of the output
variable.
d 4 y
b
a1 d 3 y a2 d 2 y a3 dy a4
=

2
y+ 4 u
4
3

dt
a0
a0 dt
a0 dt
a0 dt a0
(7.50)

(7.49)
dx4 d 4 y
= 4
dt
dt

(7.53)

and using the state variables to eliminate output variable y


and its derivatives on the right side
a
a
a
a
b
 dx4
= 1 x4 2 x3 3 x2 4 x1 + 4 u
dt
a0
a0
a0
a0
a0
(7.54)
Summarizing the state equations


dx1
dt
dx2
dt
dx3
dt
dx4
dt

= x2
= x3
= x4
a
a
a
a
b
= 1 x4 2 x3 3 x2 4 x1 + 4 u
a0
a0
a0
a0
a0

(7.55)

7.9 Expression of Systems Equations in State-Space

439

Rewriting this equation using x1 for the output variable v3 g , u


for the input variable v1g , and 1, 2, and for the constant
terms

v(t)

d x1
= x2
dt

Fig. 7.6 Linear graph of the series RLC circuit of Fig.7.5

Expressing the state equations in matrix form,

0
x1
0

d x2
0
=
dt x3
a4
x4
a0

1
0
0

0
1
0

a
3
a0

a
2
a0

0
0

0
x1 0
x

1 2 + 0 u
x3

a1 b4
x4
a0
a0

and substituting

x1
x
2
y = [1 0 0 0]
x3

x4

(7.57)

7.9.1.1Example Expression of a Higher-Order


System Equation Without Differentiation
of the Input as State Equations
We again use the series RLC circuit as the example system.
The linear graph, Fig.7.6, and the system equation for the
output variable is the voltage across the capacitor, v3g, from
Sect.7.6.2 are repeated below for convenience.
The second-order system equation is:
d 2 v3 g R dv3 g
1
1
v1g =
+
+
v3 g
LC
L dt
LC
dt 2
Rearrange the system equation so that the highest-order derivative of the output variable is on the left side. We create
a summation of three terms: the output variable, its first derivative, and the input variable, each multiplied by a factor
comprising the system parameters.
d 2 v3 g
dt

(7.60)

dx
= Ax + Bu,
dt

and the output equations in matrix form, y = Cx + Du,

(7.59)

We see that the equation can be expressed in the form of a


state equation by defining

dx1
d dx1
+ u

= 1 x1 + 2
dt dt
dt

1
1
R dv3 g
v3 g
+
v1g
L dt
LC
LC

(7.58)

(7.56)

dx2
= 1 x1 + 2 x2 + u
dt

(7.61)

These two equations are a set of state equations.


Applying this method to the system equation of the RLC
circuit,
d 2 v3 g
dt

1
1
R dv3 g
+
v3 g
v1g
LC
L dt
LC

we define the output variable as the first state variable


v3 g x1
and derivative of the output variable as the second state variable. It is essential to recognize that this definition also creates a state equation.
dv3 g
dt

x2 =

dx1
dt

This is a recursion, meaning the next variable is defined


in terms of the previous. The second derivative of the output
variable equals the derivative of the second state variable.
d 2 v3 g
dt

dx2
dt

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

440
Table 7.5 State and output eqs.
for the RLC circuit of Fig.7.6

Energy storage variables as state variables

State variables defined recursively from the output variable of the second-order system equation

1
R
L L iL 1

+ L v1g

v
1

0 3 g 0
C

iL
v3 g = [ 0 1]
v3 g

0
d x1
=
1

dt x2
LC

d iL
=
dt v3 g

Substituting into the system equation yields


dx2
1
1
R
=
x1 x2 +
v1g
dt
LC
L
LC
We now have the set of state equations
dx1
= x2
dt
dx2
1
1
R
=
x1 x2 +
v1g
dt
LC
L
LC
The output variable of the system is the voltage across the
capacitor, v3 g . We use our definition of the first state variable as an algebraic output equation to express the output in
terms of the state variables
v3 g = x1
In this particular case, the output equation is just the definition of state variable. However, in the case of systems in
which the input is differentiated, then state variable x1 and
output variable will not be the same.
Expressing state equations and the output equation in matrix form.
dx
= Ax + Bu
dt
0
d x1
=
1
dt x2
LC
y = Cx + Du

1
0
x1

+
R
1 v1g
x2
L
LC

x1
v3 g = [1 0]
x2

When we compare these state and output equations to those


we derived using the systems energy storage variables as the
state variables, we see that they are not the same, as shown
in Table7.5.

1
0
x1
R + 1 v1g

x2
L
LC

x
v3 g = [1 0] 1
x2

7.9.2Algorithm to Express a Higher-Order


System Equation with Differentiation
of the Input as State Equations
The algorithm to express higher-order system equations with
derivative input terms is slightly more involved. It is based
on the fact that state variables are not unique. We can define
any number of sets of state variables with algebra. In this
algorithm, we define a set of state variables consisting of
weighted sums of the input and output variables, where
the term weighted means each variable is multiplied by a
different coefficient.
State equations are sets of simultaneous first-order differential equations, where the only derivative in an equation
is the state variable on the left-hand side. There can be no
derivatives on the right-hand side, including derivatives of
the inputs.


dx
= Ax + Bu
dt

(7.45)

We will see when we develop the transfer function representation of higher-order systems that it is common for the input
variable to appear as one or more derivatives. Derivatives of
the inputs are dealt with by defining non-physical state variables, which are algebraic expressions, typically including
a state variable and a derivative of the input variable. This
sounds intimidating, but there is a reasonable algorithm to
handle these cases.
The algorithm assumes that both the input and output
variables are differentiated to the same degree. If this is not
the case, as it typically isnt, then eliminate unneeded input
derivatives, by setting their respective coefficients to zero.
We will use a fourth-order system equation for illustration,
where the input variable is u (t ), and the output variable is
y (t ).

d4y
d3y
d2y
dy
a 0 4 + a1 3 + a 2 2 + a 3
+ a4 y
dt
dt
dt
dt
d 4u
d 3u
d 2u
du
= b 0 4 + b1 3 + b 2 2 + b 3
+ b 4u
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


is to clear the highest-order derivative of the output variable
of any coefficient term.
 d y a 1 d y a 2 d y a 3 dy a 4
+
+
+
+ y =
dt 4 a 0 dt 3 a 0 dt 2 a 0 dt a 0
4

Define the first state variable, x1 , as the operand of the highest-order derivative.
b0
x1 = y u
a0

b 0 d 4 u b1 d 3u b 2 d 2 u b 3 du b 4
+ u
4 + 3 + 2 +
a 0 dt
a 0 dt
a 0 dt
a 0 dt a 0
(7.63)

The remaining state variables are defined in terms of the derivative of the output variable y, the input variable u, and
coefficient terms i
x1 = y 0 u
dx
x2 = 1 1u
dt
dx2
x3 =
2 u
dt
dx
x4 = 3 3u
dt

The basis of the technique is to define new state variables,


which both incorporate the derivatives of the input and allow
us to write the state equations in the same form, as the case
without derivative of the input, i.e.,


1
0
x1 0
0

d x2 0
0
=
dt x3
a3
a4
x4
a0
a 0
x1
x
2
y = [1 0 0 0] + 0u
x3

x4

0
1
0
a2

a0

0
0
1

x1 1

x2 + 2 u
x3 3
a1
x4 4
a 0

(7.64)

a 1 d 3 y a 2 d 2 y a 3 dy a 4
d4y
= 3 2
y
4
dt
a 0 dt
a 0 dt
a 0 dt a 0
b 0 d 4 u b1 d 3u b 2 d 2 u
+ 4 + 3 + 2
a 0 dt
a 0 dt
a 0 dt
b 3 du b 4
+
+ u
a 0 dt a 0

(7.65)

(7.68)

and


0 =
1 =
2 =
3 =

First, solve for the highest-order output variable on the left


side.

(7.67)

4 =

b0
a0
b1
a0
b2
a0
b3
a0
b4
a0

a1
a0
a1
a0
a1
a0
a1
a0

0
1
2
3

a2
a0
a2
a0
a2
a0

0
1
2

a3
a0
a3
a0

0
1

a4
a0

(7.69)

7.9.2.1Example Expression of a Second-Order


System Equation with Differentiation of
the Input as State Equations
We will demonstrate this method with the translational system, consisting of a force source acting on the mass attached
to a spring and a damper of Sect.3.2. The schematic and
linear graph are Fig.7.16.

Collect the input and output variable terms by order of differentiation to guide our choice of state variable.

d4
dt 4

b0 d 3 a1
b1 d 2
y u = 3 y + u + 2
a 0 dt a 0
a 0 dt

a2
b2 d a 3
b3 a 4
b
y + u + y + u y + 4 u
a 0 dt a 0
a 0 a 0
a0
a 0
(7.66)

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

442
Fig. 7.16a Schematic and b linear graph of a spring-mass-damper system with an input force

x,v

F(t)

F(t)

K
g

The second-order system equation, which relates the force


input to the velocity of the mass, Eq.3.34, has differentiation
of the input term with respect to time.
2
1 dF (t ) d v1g b dv1g K
=
+
+ v1g
M dt
M dt
M
dt 2

(3.34)

The system is of second order. Hence, it will be represented


by a set of two state equations of the form


1
0
x
d x1
a1 1 + 1 u
= a2

x
dt 2 x2 2
a0
a0
x1
y = [1 0] + 0 u
x2

a0
(7.70)

where


0 =
1 =
2 =

b0
a0
b1
a0

b2
a0

a1
a0
a1
a0

0
1

and
a2
a0

x1 = y 0 u
x2 =

Instruments calculators use the convention, which identifies


the coefficient of the zero-order term with the subscript of
zero.
You must identify which of the two conventions is used,
before applying an algorithm which identifies coefficients
with subscripts. Authors should identify which convention
they are using, but they do not always. If the convention is
not identified, then you must not use the algorithm. We will
identify the convention used by showing a differential equation with the coefficients identified, thusly:

dx1
1u
dt

(7.71)
There are two conventions for identifying the coefficients in
polynomials and differential equations. We will use the convention, which identifies the coefficient of the highest-order
term with the subscript of zero, a0. The other convention
identifies the coefficient with a subscript equal to the order
of the term, with the constant or zero-order term a0. Texas

d 2v
dv
d 2F
dF
a
a
v
b
+
+
=
+ b1
+ b2 F
1
2
0
2
2
dt
dt
dt
dt

Note that this differential equation is not in standard form,


since the standard form would have no coefficient on the
highest-order term of the output variable, v, in this case.
By convention, the coefficients ai are assigned to the output variable terms and the coefficients bi are assigned to the
input variable terms.
First, rewrite the system equation in the required form
with the input variable differentiated to the same order as
the output variable, adding terms with coefficients of zero,
if needed
d 2 v1g
dt 2

b dv1g K
d 2 F 1 dF
+ v1g = 0 2 +
+ 0F
M dt
M
M dt
dt

Next, equate the coefficients of the system equation with the


coefficients in the required form
a0 = 1
b
a1 =
M
K
a2 =
M
Evaluate Eqs 7.71.

and

b0 = 0
1
b1 =
M
b2 = 0

7.10 Eigenvalues and Eigenvectors

0 =
1 =
2 =

443

b0
a0
b1
a0

b2
a0

a1
a0
a1
a0

x1 = y 0 u
dx
x2 = 1 1u
dt

a2
a0

0 =

0
=0
1

1 =

1
b
1
0 =

M M
M

2 =

0 b
b 1
b
K
1
0 =
= 2
1 M
M M
M
M

x1 = v 0 F
dx
x2 = 1 1 F
dt

1
1
M
x

1
F
b +

x2 b
M
M 2

dx
= Ax + Bu
dt

x1
y = [1 0] + 0 F
x2

dx
= L { Ax + Bu}
dt

L ddtx = AL {x} + BL {u}

7.10.1Eigenvalues

As the algorithm for expressing a system equation in the


state-space illustrated, state variables and their corresponding state equations are not unique. We can define a new set
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
state equations which are the same. Specifically, all of the
different sets of state equations must yield the same characteristic equation, since the characteristic equation, indeed,
characterizes the physical attributes of the systems homogeneous, or natural, response to a disturbance. The roots of the
characteristic equation, the characteristic values, are more
commonly known by their combination GermanEnglish
name, eigenvalues. Eigen translates from German as
ones own or inherent, or, in other words, characteristic.
Conventionally, characteristic values are represented by the
Roman character, s, and eigenvalues by the Greek character,
, but, again, they are of the same quantity.
The algebra of creating a new set of state variables and
state equations is known as a transformation. In the abstract, a transformation is straightforward. Define a square
transformation matrix P and a new vector of state variables
z, by equating their product with the existing state vector x.
x = Pz

L ddtx = L {Ax} + L {Bu}

7.10 Eigenvalues and Eigenvectors

x1 = v
dv 1
x2 =
F

dt M

In order to perform the needed algebra, the state equation


dx
= Ax + Bu must be converted into algebraic equations by
dt
use of Laplace transformation.

Substituting into the state and output equations yields


0
d x1
=
K
dt x2
M

(7.72)

sX ( s ) = AX ( s ) + BU ( s )

(7.73)

Perform Laplace transformation on Eq.7.71. The transformation matrix is a constant matrix.

L {x} = L {Pz}

L {x} = PL {z}

X( s ) = PZ( s )
(7.74)
We can now eliminate the existing state vector, X(s), from
the state equations 7.72 by substitution.


sPZ ( s ) = APZ ( s ) + BU ( s )

(7.75)

The left side of Eq.7.74 is not in the proper form for state
equations. We must remove matrix P from the product,
sPZ(s). To do so, we premultiply both sides of Eq.7.74 by
the inverse of matrix P.
P 1 sPZ ( s ) = P 1 ( APZ ( s ) + BU ( s ))


P 1 sPZ ( s ) = P 1 APZ ( s ) + P 1BU ( s )

(7.76)

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

444

Although the Laplace variable s is complex, it is a scalar


variable and can be factored out of the product on the left
side of Eq.7.75.


sP PZ ( s ) = P APZ ( s ) + P BU ( s )
1

(7.77)

The product of the matrix P and its inverse matrix P1 is the


identity matrix I


sIZ ( s ) = P 1 APZ ( s ) + P 1BU ( s )

(7.78)

The identity matrix I can be inserted, if the resulting product


can be evaluated, or removed at will. We remove it to yield
the transformed state equations in the standard form.


sZ ( s ) = P 1 APZ ( s ) + P 1BU ( s )

(7.79)

where the products, P 1 AP and P 1B, are the transformations


of the original coefficient matrix A and input matrix B.
State-space based automatic control is known as Modern
Control, because it was modern in the 1960s. The transformation illustrated in Eqs.7.737.79 is a key step in Modern
Control design. The objective of the transformation is to use
a transformation matrix P, which yields a diagonal matrix
P 1 AP. The significance is that the elements of the diagonal
are the eigenvalues of the system


z1 1 0 0 0 z1
z 0
0 0 z2
2
2
+ P 1BU ( s ) (7.80)
s =
     


0 0 n zn
zn 0

allowing the control designer to change the eigenvalues individually, and tailor the response of the dynamic system under
feedback control.

An eigenvalue and its associated eigenvector are defined by


the scalar-matrix equation.
Av = v

a11
a
21

a12 v1
v1
=

a22 v2
v2

(7.81)

Note that the left side of the equation is the product of a matrix and a vector, and the right side is the product of a scalar
and a vector. The equality of the two sides is easiest to see,

a11v1 + a12 v2 v1
a v + a v = v
22 2
21 1
2

Just as eigenvalues have physical meaning when they are associated with a physical system, so too their corresponding
eigenvectors. In the case of a lumped parameter mechanical system, i.e., a spring-mass-damper system, with multiple
masses and an oscillatory impulse or step response, the eigenvalues represent the different decay rates and frequencies. Their corresponding eigenvectors represent the shape
of each mode superimposed to create the vibration.
In the bad old days, engineering students studied arcane
methods for calculating eigenvectors. In the present day,
Mathcad, MATLAB, and hand-held engineering calculators
have built-in routines for finding eigenvectors. Using the
coefficient matrix of the RLC circuits state-space representation, with energy storage variables as state variables, and
letting R = 3, L = 2, and C = 1, for illustration,
R
L
A=
1
C

1 3

L
2
=
1
0
1

1

2

The eigenvalues are


2

1 , 2 =

R
4
3
4
R
3


L
2
L
LC
2
2 1
=
2
2

1 = 0.5 and 2 = 1.0


The corresponding eigenvectors, calculated with Mathcad,
are
0.447
v1 =

0.894

7.10.2Eigenvectors

by expressing the equation in a form which can be expanded.


Using a two by two coefficient matrix; for example,

and

0.707
v2 =

0.707

Summary
Linear algebra is a condensed form of the algebra of simultaneous equations where the variables are represented as vectors and the constants or coefficients are represented by matrices. The key differences between scalar and vector-matrix
algebra are:

Problems

445

Vector-matrix multiplication is not commutative, except


in the two special cases below.
AB BA
Scalar operators are applied to each element with a matrix.
Scalar operations on a matrix are commutative.
aB = Ba
There is no vector-matrix division; it is not defined. The
equivalent is to multiply a matrix by its inverse, which is
the quantity that when multiplied by the matrix yields the
identity matrix I.
AA 1 = A 1 A = I
where I is the identity matrix:

y = Cx + Du

AI = IA

where y is the vector of output variables, C is the output


matrix, and D is the direct pass-through matrix.

and
1
0
I=


0 0
1  0
  

0 0 1

Problems

State-space represents a dynamic system as a set of firstorder simultaneous differential equations:


x1 a 1,1

d x2 a 2,1
=
dt  

xn a n ,1

transformation. State variables created by a transformation


usually have no physical meaning.
The reduction of the energetic equations of a system to
a set of state equations is significantly easier than the corresponding reduction to a higher-order differential equation
because more variables are retained, leaving a few that must
be eliminated by substitution.
The state equations are solved by finite difference methods, specifically the RungeKutta algorithm. The solution
yields the state variable values over the duration of the solution.
The solution for a power variable other than a state variable requires a second set of purely algebraic equations
called the output equations. The output equations cannot
be solved until the state equations have been solved and the
state vector x is known. The output equations in vector-matrix form are:

a 1,2
a 2,2

a n ,2

Problems 1, 2, and 3 are to be solved manually, meaning, that you may use your calculator to perform arithmetic
operations but not linear algebra. After you have solved the
problems manually, you are encouraged to check your man-

 a 1, n x1 b1,1 b1,2
 a 2, n x2 b 2,1 b 2,2
+
   


 a n , n xn b n ,1 b n ,2

Expressed in vector-matrix form as:


dx
= Ax + Bu
dt
where x is the vector of state variables, known as the state
vector, A is the coefficient matrix, B is the input matrix, and
u is a vector of inputs.
The state variables used to reduce the energetic equations
of a system, extracted from a schematic or a linear graph, are
the energy storage variables. However, the state variables of
a system are not unique since linear algebra performed on
the state equations can be used to redefine the state variables
as a combination of a previous set, in what is known as a

 b1, m u 1 (t )

 b 2, m u 2 (t )
  

 b n , m u m (t )

ual linear algebra computations using your calculator so that


you become familiar with its capabilities.
Problem 7.1 Evaluate the following matrix expressions
manually
7.1.a

1 2 5 6
3 4 + 7 8 =

1 2 5 6
7.1.b 8

=
3 4 7 8

Problem 7.2 Evaluate the following matrix expressions


manually
7.2.a

1 2 5 6
3 4 7 8 =

5 6 1 2
7.2.b

=
7 8 3 4

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

446

x,v

F(t)

Flywheel
Rotational Inertia J

Compliant Shaft
Torsional Spring K
Fluid Coupling
Damping b

Lubricating fluid
Damping b

Angular Velocity
Input (t)

Fig. P7.4 Second-order translational mechanical system


Fig. P7.5 Second-order rotational mechanical system

1 2 5
7.2.c
=
3 4 6

7.2.d

[3

5
4] =
6

5
7.2.e [3 4] =
6
Problem 7.3 Calculate the inverses of the following matrices
manually, as the transpose of the cofactors of the matrix divided by the determinant of matrix. No credit will be awarded
for inverting the matrices using a graphical algorithm. Check
by multiplying the matrix by its inverse manually.
1

7.3.a

2 3
4 5 =

2 0 3
7.3.b 4 5 6
7 8 9

undamped natural frequency n, the damped natural


frequency d, and the damping ratio .
7.4.f The system is de-energized when it is acted on by
a step input of 10N. Solve the state equations and
the output equations using Mathcad or MATLAB for
Cases I and II. Plot the responses of the output variables.
Problem 7.5 A rotational mechanical system consisting of
a spring, an inertia, and a damper acted upon by an applied
torque is shown in the schematic Fig.P7.5.
Table P7.5 Parameter values
J

Case I

Case II

Problem 7.4 A translational mechanical system consisting


of a mass M sliding on a lubricating fluid film with damping
b. A spring K is attached between the mass and ground. The
mass is acted upon by an applied force F(t), Fig. P7.4.
Table P7.4 Parameter values
M

Case I

Case II

7.4.a Derive the second-order system equation for the force


acting through the spring and express it in standard
form.
7.4.b Derive the state equations for this system.
7.4.c Derive the output equations for:
i The velocity of the mass M.
ii The force acting through spring K.
iii The angular velocity difference across spring K.
7.4.d Express the state and output equations in matrix form.
7.4.e For Cases I and II, calculate the systems eigenvalues.
If the system is underdamped, determine the ideal,

7.5.a Derive the second-order system equation for the


torque in the spring and express it in the standard
form.
7.5.b Derive the state equations for this system.
7.5.c Derive the output equations for:
i The angular velocity of the inertia J.
ii The torque acting through the compliant shaft
spring K.
iii The angular velocity difference across the compliant shaft spring K.
iv The angular velocity difference across the fluid
coupling b.
7.5.d Express the state and output equations in matrix form.
7.5.e For Cases I and II, calculate the systems eigenvalues.
If the system is underdamped, determine the ideal,
undamped natural frequency n, the damped natural
frequency d, and the damping ratio .
7.5.f The system is de-energized when it is acted on by a
step input of 10rad/sec. Solve the state equations and
the output equations using Mathcad or MATLAB for
Cases I and II. Plot the responses of the output variables.

Problems

447
x,v

K
F(t)

b2

b1

M1

Problem 7.7 A fluid system consisting of a fluid power unit


modeled as a pressure source, three fluid resistances, a fluid
accumulator with capacitance C, a fluid inertance I, and a hydraulic piston/cylinder driving a mass M is shown is shown
in the schematic, Fig.P7.7. Note that the mass and the fluid
inertance are dependent energy storage elements.

M2

Table P7.7 Parameter values

Fig. P7.6 Third-order mechanical system

Problem 7.6 A translational mechanical system consisting


of a spring, two masses, and two dampers acted upon by an
applied force is shown in the schematic Fig.P7.6.
Table P7.6 Parameter values
M1

M2

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

7.6.a Derive the state equations for this system.


7.6.b Derive the output equations
i The velocity of mass M1.
ii The force acting through the spring K.
iii The velocity of mass M2.
iv The force acting through damper b1.
v The force acting through damper b2.
7.6.c Express the state and output equations in matrix form.
7.6.d Calculate the systems eigenvalues for Cases I and
II. If the system is underdamped, determine the ideal,
undamped natural frequency n, the damped natural
frequency d, and the damping ratio .
7.6.e The system was de-energized when a step input of
10 was applied. For Cases I and II, solve the state
equations and the output equations using Mathcad or
MATLAB. Plot the responses of the output variables.

R1

R2

R3

0.1

0.2

0.3

7.7.a Derive the state equations for this system and check
their units
7.7.b Derive output equations for:
i The pressure in the fluid accumulator.
ii The velocity of mass M.
iii The volume flow rate through the fluid inertance.
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.
7.7.c Express the state and output equations in matrix form.
7.7.d Calculate the systems eigenvalues. If the system is
underdamped, determine the ideal, undamped natural
frequency n, the damped natural frequency d, and
the damping ratio .
7.7.e The system was de-energized when a step input of 100
was applied. Solve the state equations and the output equations using Mathcad or MATLAB. Plot the
responses of the output variables.
Problem 7.8 A hybrid fluid-translational rotational system
is shown in Fig.P7.8. The fluid power unit is modeled as a
pressure source, which discharges into a hydraulic line with
fluid resistance R. The hydraulic piston/cylinder has diameter D. Its piston rod is attached to the linkage. A dashpot
with damping b and a spring K are also attached to the linkage. The linkage has rotational inertia J.

Fig. P7.7 Hybrid fluid-translational mechanical system

Fluid Power Unit


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

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

448
Fig. P7.8 Fluid-translational
mechanical system

Fluid Power Unit


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
Rotational Inertia J

Fig. P7.9 Translational mechanical system

Mass M

Lubricating Film
Damping b 2

Pivot

Dashpot b1

Spring K

Spring K2

Spring K1

L3
L2

Force Source F(t)

Table P7.8 Parameter values


R

L1

L2

0.6

0.12

50

20

1,500

7.8.a Derive the state equations for this system and check
their units.
7.8.b Derive output equations for:
i The volume flow rate from the source.
ii The force applied by the piston rod.
iii The force in spring K.
iv The angular velocity of the linkage.
v The velocity difference across the spring.
7.8.c Express the state and output equations in matrix form.
7.8.d Calculate the systems eigenvalues. If the system is
underdamped, determine the ideal, undamped natural
frequency n, the damped natural frequency d, and
the damping ratio .
7.8.e The system was de-energized when a step input of
2,000 was applied. Solve the state equations and the
output equations using Mathcad or MATLAB. Plot the
responses of the output variables.

L1

Problem 7.9 A translational mechanical system is shown


schematically in Fig.P7.9. The system consists of a force
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
M, which slides on a rail with a lubricating film, damping
b2. The system is de-energized before the input step force of
1,000N acts on the system.
Table P7.9 Parameter values
L1

L2

L3

b1

b2

K1

K2

100

0.5

1,000

1,500

10

7.9.a Derive the state equations for this system and check
their units.
7.9.b Derive output equations for:
i The force acting to accelerate mass M.
ii The velocity of mass M.
iii The force acting through spring K1.
iv The force acting through spring K2.
v The velocity of the point of application of the
input force F(t).

Problems

449

Fig. P7.10 Hybrid electromechanical system

Electric Motor
Transduction

Rotational Inertia J 1
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.


7.9.d Calculate the systems eigenvalues. If the system is
underdamped, determine the ideal, undamped natural
frequency n, the damped natural frequency d, and
the damping ratio .
7.9.e Solve the state equations and the output equations
using Mathcad or MATLAB. Plot the responses of the
output variables.
Problem 7.10 An electromechanical system is shown
in Fig.P7.10. The input to the system is a voltage which
drives a DC motor through wires with electrical resistance
R. The relationship between the motors current and torque
is TM= iM. Flywheel J1 is attached to the motors shaft and
is supported by bearings with damping b1. Coupled to the
motors shaft is a compliant shaft modeled as torsion spring
K. The compliant shaft drives flywheel J2,which is supported
by bearings with damping b2.
Table P7.10 Parameter values
R

J1

b1

J2

b2

0.6

1,000

0.5

7.10.a Derive the state equations for this system and check
their units.
7.10.b Derive output equations for:
i The current from the source.
ii The torque acting to accelerate angular velocity
of rotational inertia J1.
iii The angular velocity of rotational inertia J1.
iv The angular velocity of rotational inertia J2.
v The torque acting through spring K.
7.10.c Express the state and output equations in matrix
form.
7.10.d Calculate the systems eigenvalues. If the system is
underdamped, determine the ideal, undamped natural
frequency n, the damped natural frequency d, and
the damping ratio .

7.10.e The system was de-energized before a step input of


48 VDC was applied. Solve the state equations and
the output equations using Mathcad or MATLAB.
Plot the responses of the output variables.
Problem 7.11 A rotational system show is shown in Fig.7.11.
An angular velocity source drives the shaft attached to the
pulley with diameter D1. The belt over pulley D1 drives the
pulley with diameter D2, which is attached to the same shaft
as the pulley with diameter D3. That shaft is compliant, with
spring constant K, and drives the flywheel with rotational
inertia J1 and damping b1. The belt over pulley D3 drives the
pulley with diameter D4, which is attached to a shaft connect
to the fluid coupling with damping b2. The output shaft of the
fluid coupling drives the flywheel with rotational inertiaJ2
and damping b1.
Table P7.11a Parameter values
D1

D2

D3

D4

b1

J1

J2

b2

1,500

12

Table 7.11b Torsion spring stiffness K values


Case I

Case II

Case III

500

1,500

3,000

7.11.a Derive the state equations for this system and check
their units.
7.11.b Derive output equations for:
i The torque acting through spring K.
ii The angular velocity of flywheel J1.
iii The torque applied by the angular velocity source.
iv The torque acting to accelerate flywheel J2.
v The angular velocity of flywheel J2.

450

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

Fig. P.7.11 Rotational system

Flywheel
Rotational Inertia J1
Damping b1

Compliant Shaft
Spring Constant K

Pulley
Diameter D1

Flywheel
Rotational Inertia J 2
Damping b1

Pulleys Diameters
D2 and D3

Angular Velocity
Input (t)

Fluid Coupling
Damping b2

Pulley
Diameter D4
Fig. P7.12a A shaft-driven rotational system with a compound
gear

Rotational Inertia J
Compliant Shaft
Torsion Spring K2
Compliant Shaft
Torsion Spring K 1

Bearing
Damping b3
N2

N4
N3

Bearing
Damping b2

N1

Fluid Coupling
Damping b1
Angular Velocity
Source, (t)

7.11.c Express the state and output equations in matrix


form.
7.11.d Calculate the systems eigenvalues for Cases I,
II, and III using the damping coefficient values of
Table 7.11b. If the system is underdamped, determine the ideal, undamped natural frequency n,
the damped natural frequency d, and the damping
ratio.
7.11.e The system was de-energized before a step input of
1,500 RPM was applied. Solve the state equations
using Mathcad or MATLAB.
7.11.f Solve and plot the responses of the output variables
using Mathcad or MATLAB.
Problem 7.12 A rotational mechanical system is shown
schematically in Fig. P7.12. The system consists of an angular velocity source (t) acting on the input shaft of a fluid
coupling with damping b1. The output shaft of the fluid coupling drives inertia J1, which drives a compliant shaft with

torsional spring constant K. The compliant shaft drives a pinion with N1 teeth. The pinion engages a gear with N2 teeth.
The gear shaft drives rotational inertia J2. The gear shaft
bearings have damping b2.
Table 7.12a Parameter values
b2
b3
K1
20

N m sec
rad

10

N m sec
rad

2,000

K2
N
rad

4,000

J
N
rad

10 kg m 2

Table 7.12b Fluid coupling damping b1 values


Case I

Case II

Case III

N m sec
5
rad

N m sec
1
rad

0.1

N m sec
rad

Table 7.12c Gear teeth numbers


N1

N2

N3

N4

12

48

18

54

Problems

451

3,000

p(t)
psi

200
100

2,000

T(t)
N m

1,000

t, sec

-100
-200

t, sec

Fig. P7.13b Torque input

Fig. P7.12b Angular velocity input

7.12.a Derive the state equations and check their units.


7.12.b Derive output equations for the following variables:
i The angular velocity of the pinion.
ii The torque acting through the compliant shaft.
iii The torque acting to accelerate inertia J2.
iv The torque acting through the fluid coupling b1.
7.12.c Write the state equations and output equations in
vector-matrix form.
7.12.d Calculate the systems eigenvalues for Cases I, II,
and III. If the system is underdamped, determine the
ideal, undamped natural frequency n, the damped
natural frequency d, and the damping ratio .

The system is de-energized before the input pulse of
angular velocity shown in Fig.7.12b is applied.
7.12.e Solve and plot state equations for the velocity pulse
shown in Fig.P7.12b.
7.12.f Solve and plot the output equations.
Problem 7.13 A rotational-translational mechanical system
is shown schematically in Fig. P7.13. The system consists
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 rotationaltranslational system


Gear
N 2 teeth
Inertia J

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
power screw threads through the crosshead. The crosshead
translates when the power screw rotates. The viscous friction
between the rotating power screw and the translating crosshead is modeled as rotational damping b1. The crosshead has
mass M and slides on two rails. The combined viscous friction of the crosshead sliding on the two rails is modeled as
translational damping b2.
7.13.a Derive the state equations and check their units.
7.13.b Derive output equations for the following variables:
i The angular velocity of the pinion shaft.
ii The angular velocity gear N2.
iii The torque in the compliant shaft.
iv The force acting to accelerate the crosshead.
v The translational velocity of the crosshead.
7.13.c Write the state equations and output equations in
vector-matrix form.
7.13.d Calculate the systems eigenvalues. If the system is
underdamped, determine the ideal, undamped natural
frequency n, the damped natural frequency d, and
the damping ratio .

Compliant Shaft
Spring Constant K

Crosshead
Mass M

Rail
Power Screw
4 threads per inch

T(t)
Pinion
N1 teeth

Viscous Friction of
threads in crosshead
Rotational Damping b1

Viscous Friction of crosshead


sliding on rail Damping b2
(total for both rails)

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

452
Fig. P7.14a Hybrid fluid-rotational mechanical system

Flywheel, Inertia J and Damping b


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


N1 N2 J
b1
12

48

1 kg m 2

N m sec
rad

2,000

N
rad

b2

5 kg

0.2

N m sec
rad

The system is de-energized before the input torque


pulse shown in Fig.P7.13b was applied to the system.
7.13.e Solve the state equations using Mathcad or MATLAB for the input torque pulse shown in Fig.P7.13b
with the parameter values of Table7.
7.13.f Solve and plot the responses of the output variables
using Mathcad or MATLAB.
Problem 7.14 A fluid-rotational mechanical system is shown
in Fig.P7.14. The fluid power unit is modeled as pressure
source p(t), drawing fluid from a reservoir vented to the
atmosphere. The pumps high-pressure line has fluid resistance R and inertance I. The rotational hydraulic motor produces 600 ft.-lbs. of torque at a hydraulic fluid pressure of
3,000 psi. The fluid discharging from the motor returns to
the reservoir. The output shaft of the motor turns a compliant shaft with torsion spring constant K. The compliant shaft
drives a flywheel with mass moment of inertia J supported
on bearings with damping b.
Table 7.14 Parameter values
R
I
b

MPa sec
80
m3

N
4,000
rad

5 kg m 2

kg
24,000 4
m

N m sec
10
rad

p(t)
psi

3,000
2,000
1,500
1,000

t, sec

Fig. P7.14b Pressure input

7.14.a Derive the state equations and check their units.


7.14.b Derive output equations for the following variables:
i The volume flow rate from the fluid power unit.
ii The pressure in the fluid accumulator.
iii The angular velocity of the hydraulic motor.
iv The torque acting to accelerate the flywheel.
v The angular velocity of the flywheel.
7.14.c Write the state equations and output equations in
vector-matrix form.
7.14.d Calculate the systems eigenvalues. If the system is
underdamped, determine the ideal, undamped natural frequency n, the damped natural frequency d,
and the damping ratio . If the real components of
the eigenvalues vary by two or more orders of magnitude, then the duration of the components of the
response vary by like magnitude. If so, solve the
state equations twice, once using a short duration to
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 mechanical system

Shaft rigidly attached


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 (t)

7.14.e Solve the state equations using Mathcad or MATLAB


for the input pressure pulse shown in Fig.P7.13b
with the parameter values of Table7.14.
7.14.f Plot the responses of the output variables using
Mathcad or MATLAB.
Problem 7.15 A rotational system is shown in Fig.7.15. An
angular velocity source drives the input shaft to a fluid coupling with damping b1. The output shaft of the fluid coupling
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 flywheel J2 with rotational inertia J2. A drag cup with damping
b2 is connected between the hub of flywheel J2 and the machine frame, which is ground.
Table 7.15a Parameter values
K1
600

Nm
rad

J1

K2

2 kg m 2

400

Nm
rad

J2

b2

1 kg m 2

0.2

N m sec
rad

Table 7.15b Fluid coupling damping b1 values


Case I
2

N m sec
rad

Case II
20

N m sec
rad

Case III
40

N m sec
rad

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.
ii The torque acting through the compliant shaft.
iii The torque acting to accelerate inertia J2.
iv The torque acting through the fluid coupling b1.
7.15.c Write the state equations and output equations in
vector-matrix form.
7.15.d Calculate the systems eigenvalues for Cases I, II,
and III. If the system is underdamped, determine the
ideal, undamped natural frequency n, the damped
natural frequency d, and the damping ratio .

L2

L1
v(t)

C1

C2

Fig. P7.16 An RLC circuit

7.15.e The system was de-energized before a step input of


1,500 RPM was applied. Solve the state equations
and plot their responses using Mathcad or MATLAB
for cases I, II, and III.
7.15.f Solve and plot the responses of the output variables
using Mathcad or MATLAB for cases I, II, and III.
7.15.g Calculate and plot the power dissipated in the fluid
coupling b1 and the drag cup b2 for cases I, II, and III.
Problem 7.16 The schematic of an electric circuit with a
voltage source, two inductors, two capacitors and a resistor
is shown in Fig.P7.16.
Table 7.16a Parameter values
C1

C2

L1

L2

1 F

2 F

1 mH

2 mH

Table 7.16bResistor R values


Case I

Case II

Case III

20

40

7.16.a Derive the state equations and check their units.


7.16.b Derive output equations for the following variables:
i The current flowing through capacitor C1.
ii The current flowing through capacitor C2.

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

454

L1
v(t)

Table 7.17bResistor R values

L2

C1

C2

iii The voltage across inductor L1.


iv The current through resistor R.
7.16.c Write the state equations and output equations in
vector-matrix form.
7.16.d Calculate the systems eigenvalues for Cases I, II,
and III. If the system is underdamped, determine the
ideal, undamped natural frequency n, the damped
natural frequency d, and the damping ratio .
7.16.e The system was de-energized before a step input of
24 VDC was applied. Solve the state equations and
plot their responses using Mathcad or MATLAB for
cases I, II, and III.
7.16.f Solve and plot the responses of the output variables
using Mathcad or MATLAB for cases I, II, and III.
7.16.g Calculate and plot the power dissipated in resistor R
for cases I, II, and III.

Table 7.17a Parameter values


L1

L2

1 F

2 F

1 mH

2 mH

Fig. P7.18a Electromechanical


system with a compliant shaft
and a compliant belt

Case III

10

20

Problem 7.18 An electromechanical system is shown in


Fig.P7.18a. A voltage source drives a DC motor, which has
resistance R and torque constant KT. The DC motor turns a
compliant shaft, modeled as a torsion spring with spring constant K1.
The shaft turns flywheel1 with mass moment of inertia J1.
Flywheel1 is supported on bearings with damping b1. Flywheel2 has mass moment of inertia J2 bearings with damping
b2, and is belt driven. The belt which runs between flywheels
one and two is long enough that the energy stored in the taut

Problem 7.17 The schematic of an electric circuit with a


voltage source, two inductors, two capacitors and a resistor
is shown in Fig.P7.17.

C2

Case II

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.
ii The current flowing through capacitor C2.
iii The voltage across inductor L1.
iv The current through resistor R.
7.17.c Write the state equations and output equations in
vector-matrix form.
7.17.d Calculate the systems eigenvalues for Cases I, II,
and III. If the system is underdamped, determine the
ideal, undamped natural frequency n, the damped
natural frequency d, and the damping ratio .
7.17.e The system was de-energized before a step input of
24 VDC was applied. Solve the state equations and
plot their responses using Mathcad or MATLAB for
cases I, II, and III.
7.17.f Solve and plot the responses of the output variables
using Mathcad or MATLAB for cases I, II, and III.
7.17.g Calculate and plot the power dissipated in resistor R
for cases I, II, and III.

Fig. P7.17 An RLC circuit

C1

Case I

Voltage Source v(t)

Electrical Resistance R

Compliant Shaft
Torsional Spring K1
Flywheel 1
Rotational Inertia J 1
Bearings Damping b1

DC Electric Motor
Torque constant K T
Flywheel 2
Rotational Inertia J2
Bearings Damping b 2

6 in. Diameter Pulley


Compliant Belt
Translational Spring Constant K2
4 in. Diameter Pulley

Problems

455

Fig. P7.18b The compliant


belts taut and slack sides switch
positions during an oscillation

vsurface

1g

K2

n+2

g
Fig. P7.18c Linear graph representing a compliant belt as a translational spring between two rotational to translational transducer interfaces

or stretched side of the belt, Fig.P7.18b, must be included


in the dynamic system. The compliant belt is represented
by a translational spring between two transducers, which
interface between the torque on the pulleys shafts and the
force carried by the belt, Fig.P7.18c. The sign reversal in the
translational spring represents the taut and slack sides of the
belts switching positions.
Table 7.18 Parameter values
R
KT
K1
Nm
A

K2

J2

b2

5 kg m 2

8,000

N
m

800

Nm
rad

N m sec
rad

2g

Driven
Pulley

vsurface

n+1

Taut

Taut

Driving
Pulley

n-1

vsurface

Slack

J1

b1

6 kg m 2

N m sec
rad

1g

Slack

Driving
Pulley

vsurface

2g

Driven
Pulley

7.18.a Derive the state equations and check their units.


7.18.b Derive output equations for the following variables:
i The current drawn from the voltage source.
ii The back-EMF of the DC motor.
iii The torque acting to accelerate flywheel J1.
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 systems eigenvalues. If the system is
underdamped, determine the ideal, undamped natural
frequency n, the damped natural frequency d, and
the damping ratio .
7.18.e The system was de-energized before a step input of
48 VDC was applied. Solve the state equations and
plot their responses using Mathcad or MATLAB.
7.18.f Solve and plot the responses of the output variables
using Mathcad or MATLAB.
Problem 7.19 A hybrid rotational-translational mechanical
system is shown schematically in Fig.P7.16. The system
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
has mass moment of inertia J. Spring K2 is attached to mass
M, which slides on a rail with a lubricating film, damping
b2. The system is de-energized before the input step force of
1,000N acts on the system.
Table P7.19 Parameter values
L1
L2
L3
b1
b2

K1

K2

50

1,000

1,500

10

20

Fig. P7.19 Hybrid rotationaltranslational mechanical system

Mass M
Lubricating Film
Damping b 2

Spring K2

Rotational Inertia J
Dashpot b1

Pivot

Spring K1

L3
L2

Force Source F(t)

L1

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

456
Fig. 5.52 Thermal system
divides into five regions, each
with a thermal capacitance
between the center temperature
node of the region and ground
temperature

Perfect Insulation
No heat flow

q in

R ss1

Aluminum

0.5 in

2 in

Rss 2

R al 1

Css

1(t)

R al2

RAl

RAl

Stainless
Steel

Surface Area
A = 48 ft 2

RAl RAl

RAl

RAl

RAl
3

Heat Flow
Rate

Fig. 5.53 Linear graph of the


stainless steel and aluminum
thermal system. The linear graph
is split between nodes 6 and 7 for
this figure

RSS

RSS

R al 3

RAl

11
qout

10

C al1

Ral 4

Cal 2

g
6

Ral 4

Cal 2

R al 5

Ral 6

Cal 3

R al 7

10

Ral 8

11

Cal 4

2(t)
g

7.19.a Derive the state equations for this system and check
their units.
7.19.b Derive output equations for
i The force acting to accelerate mass M.
ii The velocity of mass M.
iii The force acting through spring K1.
iv The force acting through spring K2.
v The angular velocity of the lever.
vi The velocity of the point of application of the
input force F(t).
7.19.c Express the state and output equations in matrix
form.
7.19.d Calculate the systems eigenvalues. If the system is
underdamped, determine the ideal, undamped natural
frequency n, the damped natural frequency d, and
the damping ratio .
7.19.e Solve the state equations and the output equations
using Mathcad or MATLAB.
7.19.f Plot the responses of the output variables using
Mathcad or MATLAB.
Problem 7.20 The stainless steel and aluminum thermal system of Sect.5.8.4 is reproduced as Fig.5.52. The stainless
steel is 0.5in. thick. The aluminum is 2in. thick. Both metals
are divided into layers 0.25in. thick, parallel to the external

surfaces. The thermal system model consists of thermal resistances in series and thermal capacitances in parallel. The
linear graph is Fig.5.53. The thermal resistances of the stainless steel and aluminum are calculated for 0.25in. layer. The
thermal capacitances are calculated pairing adjacent layers
and using the temperature node between them as the temperature of the 0.5 in. thick capacitance.
The thermal resistances of the 0.25in. thick stainless steel
and aluminum layers are:
RSS = 8.23 10 5

o
K
K
and RAl = 6.91 10 6
W
W

The thermal capacitances of the stainless steel and aluminum


are:
C pSS = 204

kJ
kJ
and C pAl = 137 o
K
K

7.20.a Derive the state equations for this system and check
their units.
7.20.b Derive output equations for:
i The temperatures of nodes 2, 4, 6, 8, and 10.
ii The heat flow rate through thermal resistances
RSS1, Ral1, Ral3, Ral5, Ral7, Ral9, and Ral10.

Chapter 7 Appendix

457

Fig. 3.13a Translational mechanical mass-damper system


driven by a force source, bLinear graph of the mass-damper
system driven by a force source

x,v
F(t)

Mass
M
1

Lubricating
fluid film
Damping b

Chapter 7 Appendix
Mathcads RungeKutta Solver
Mathcad has two versions of the RungeKutta algorithm,
rkfixed(), which has a constant, or fixed, time step t and
Rkadapt(). Rkadapt() adapts or varies the time step, reducing
its size if the finite differences become large and increasing
it when they are small. We will use rkfixed(). Rkadapt() is
used when a calculation is numerically stable only with an
excessive number of time steps.
The RungeKutta solver rkfixed() is a function. It is used
in an assignment statement where the result of the Runge
Kutta calculation is assigned to a variable, say Z. The
Mathcad syntax is:
Z : = rkfixed(a, b, c, d, e)
where
a is the state variable vector.
b is the beginning time of the simulation.
c is the end time of the simulation.
d is the number of time intervals in the simulation.
e is the vector containing the state equations.
The time step t=(c-b)/d.
The function rkfixed() returns an array, Z in this example.
The first column of the array is time. The succeeding columns are the values of the state variables at those times, i.e.,
the second column is the first state variable, the third column
is the second state variable, etc. A column of a matrix can be

F(t)

g
7.20.c Express the state and output equations in matrix
form.
7.20.d Calculate the systems eigenvalues.

The thermal system is at the uniform temperature of
20C when the temperature of node 1 is given a step
increase to 100C. The temperature of node 11 is
held at 20C.
7.20.e Solve the state equations and the output equations
using Mathcad or MATLAB.
7.20.f Plot the responses of the output variables using
Mathcad or MATLAB.

extracted using a special operator. Typing Ctrl and 6 simultaneously immediately following a matrix variable produces
a pair of exponentiated angle brackets. The column number
to be extracted is inserted into the angle brackets. Remember
that Mathcads origin, the beginning value of indices, is
zero, unless it is changed by the user. Hence, the first column
of a matrix is column zero. The following command assigns
the first column of the array Z to the variable t.
t:= Z0
The RungeKutta method will fail when simulating secondorder or higher systems if the time step is too large. The calculation will become numerically unstable and produce a
result with growing oscillations that head to either positive
or negative infinity. Errors formulating the state equations,
particularly sign errors, can also lead to numerical instability.
The physical systems which are not under feedback control
are physically stable. Therefore, if the calculation is unstable
reduce the time step by an order of magnitude or two. If the
calculation is still numerically unstable, then check your
state equations.
The procedure will be illustrated with a first-order massdamper system. The first-order system will also be used to
demonstrate the response of the system to inputs other than
step inputs. Next, the response of a second-order springmass-damper system will be investigated.

Step Response of a Linear Mass-Damper System


Figure 3.13 showing the schematic and linear graph of a
force source sliding a mass on a lubricating film and the
system equation for the velocity of the mass, Eq.3.26, are
reproduced here for reference.


F (t ) M dv1g
=
+ v1g
b
b dt

F (t ) dv1g b
=
v1g (3.26)
M
dt
M

Standard form of a first-order system equation requires little


rearrangement to form into a state equation.

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

458

dv1g
F (t ) dv1g b
b
F (t )
=
+ v1g
= v1g +
M
dt
M
dt
M
M
The Mathcad code begins by defining the input function, parameter values, and the initial value of the state variable. We
will first solve the state equation for a step input with the system initially de-energized using the following parameters:
M : = 3 b : = 2 F : = 10
Define a state variable vector by assigning the initial value
to each state variable. Here, the mathematical meaning of
vector collides with the computer science meaning. Each
individual state variable must be a computer science vector to contain the value of the variable at each time step in
the calculation. Recall that the period indicates a literal
subscript that Mathcad accepts as part of the variable name
[indicates what follows to be a vector or matrix subscript].
Type x[0

to see

x0

Mathcad recognizes x as a vector variable because it has vector a subscript. Define the initial value of the state variable
x to be zero.
x0 : = 0
The state equation must be defined as a function to two variables. The first argument, t, is the independent variable. The
second argument is the name of the state vector. We have a
single state variable, not a state variable, so we identify the
state variable. One can name the function which defines the
state equations. We will use D(t, x).
D ( t, x ) : =

b
F
x +
M
M

In general, the function D(t, x) contains a vector of the state


equations. In this case, there is only one equation. Notice
that the independent variable t does not need to appear on
the right side of the assignment statement. It can be used, as
demonstrated below.
First-order system difference equations cannot be numerically unstable. We will use a large number of steps, say
1,000, which is far more than is needed for accuracy.
The rkfixed function assigning its result to Z is:
Z := rkfixed(x, 0,10,1000, D)
The arguments of rkfixed() are:
x = vector of state variables.
0 = start time
10 = end time

4
V.1g
2

t1

10
10

Fig. A7.1 Step response of a linear mass-damper system

100 = number of increments


D = vector of state equations
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
when there is more than one. It is best to copy the columns
of the array to vectors.
t1: = Z 0

v1g : = Z 1

The subscripts 1g are literal subscripts produced by typing a


period after v, v.1g
This procedure hardly saves time when solving a linear
first-order equation subjected to a step input, since there are
only four possible solutions. Fortunately, we can use the
same procedure to solve some non-linear equations and all
linear equations subjected to arbitrary inputs.
First, lets compare the above response for a mass-damper
with linear damping Fb = bv1g with the response of a system
with non-linear damping FbNL = bv12g .

Step Response of a Mass-Damper System with a


Non-Linear Damper
The Mathcad code is identical, except the state equation.
We will again define the parameter values, the magnitude
of the input force, and the initial value of the state variable
so that the code is complete. There is no need to repeat these
Mathcad statements if they are unchanged.
M: = 3

b: = 2

F : = 10

Define the initial value of the state variable to be zero.


x0 : = 0
Define the state equation. Note that the state variable is
squared.

Chapter 7 Appendix

459

4
4
4

v.1g

v.1g

v.NL

( )

F t3q

2
0
4

1.357
0

t2

10
10

Fig. A7.2 Step response of a linear mass-damper system, solid trace,


and a non-linear mass-damper system, dotted trace

D ( t, x ) : =

b 2 F
x +
M
M

Assign the result of the RungeKutta function to the variable


ZNL, where the subscript NL is a literal subscript.
Z NL : = rkfixed(x, 0,10,1000, D)
It is best to copy the columns of the matrix Z NL to vectors.
We could use the same name for the time vector since they
are identical. We will create a new time vector.
t2 : = Z NL 0

v NL : = Z NL 1

Response of a Linear Mass-Damper System


Subjected to a Pulse Train
D(t, x) is a function of t, allowing the input functions of the
state equations to functions of time. Example: Fpulses(t) is a
pulse train made of step functions.
F(t) : = 5 (t) 10 (t 2) + 5 (t 4)
The remaining Mathcad code is unchanged except that the
input force in the state equations is now a function of time.
M: = 3

b: = 2

x0 : = 0
D ( t, x ) : =

b
1
x + F(t)
M
M

Assign the result of the RungeKutta function to the variable


Zpulses.
Zpulses : = rkfixed(x, 0,10,1000, D)

6
10

t3 , t3q

10

Fig. A7.3 Pulse response of a linear mass-damper system, solid trace,


and the pulse input, dotted trace

Copy the columns of the result to vectors.


t3 : = Zpulses 0

v pulses : = Zpulses 1

In order to plot the input pulse function, we define a range


variable, q, of the same length as the time vector t3. The variable q must start at zero, so we must subtract one from the
length of the vector t3.
q : = 0..length(t3) 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 Mass-Damper System to Sinusoidal


Inputs
Calculation of the response of the first-order mass-damper
system to sinusoidal inputs is a second example of the use of
the independent variable t to define a function of time as the
input to the state equations.
We will again define values even though they have been
previously defined and are unchanged so the each block of
Mathcad code is complete.
M: = 3

b: = 2

x0 : = 0
D(t, x) : =

b
1
x + sin( t)
M
M

Zpi : = rkfixed(x, 0,10,1000, D)


t4 : = Zpi 0

v pi : = Zpi 1

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

460
0.091

0.1

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

v.2pi
v.4pi

where the subscripts are vector subscripts created by typing.


The state vector is then defined as:

0.05
0.053

0.1

x 20 : = 0

x10
x:=
x 20
0

t1 , t2

10
10

Fig. A7.4 Sinusoidal responses of the first-order mass-damper system. Solid trace input frequency =2. Dashed trace, input frequency
=4

Second block of code:

v1g 0 : = 0

M: = 3 b: = 2
x0 : = 0
b
1
x + sin(4 t)
D(t, x) : =
M
M
Z4pi : = rkfixed(x , 0,10,1000, D)
t4 : = Z4pi 0

This syntax is very specific. The state vector x cannot have a


vector subscript. The state variables x1 and x2 must have the
vector subscript zero.
The alternative is to give the individual state variables
meaningful names:
FK 0 : = 0

v 4pi : = Z 4pi 1

The Mathcad code to solve higher-order state space systems


has an additional assignment statement to create a vector of
the state variables. We will illustrate the calculation using the
spring-mass-damper system, Fig.A7.5.
The first Mathcad instructions are identical to those of the
first-order mass-damper system, except that there is now a
spring constant K.
b : = 0.2

K: =1

F : = 10

Assume that the system is at rest and relaxed. Define a vector


of the initial conditions of the state variables. It is essential
that the state variables are identified in the Mathcad worksheet. A common error is to not identify which state variable
is which and then forget and use the wrong state variable in a
subsequent calculation. The state variables can be identified
Fig. A7. 5a force source-springmass-damper system. b Linear
graph of the force source-springmass-damper system

FK 0 : = 0

where the latter have literal subscripts for 1g and K and vector subscripts for zero.
The state vector x is the vector of the initial values of the
state variables where, again, the state vector x cannot have a
vector subscript and the state variables v1g and FK must have
the vector subscript of zero.

Step Response of a Spring-Mass-Damper System

M: = 3

v1g : = 0

or

v1g 0
x:=

FK
0
The state equations are:
b
D(t, x) : = M

Finite difference computations of higher-order systems


can go unstable if the time step t is too large. As a rule of
thumb, if the eigenvalues of complex conjugates, 200 steps
per period of the (of the smallest period in the system if four
order) is sufficient for a stable and accurate calculation. If
the plots of the state equations blow up (head off to positive
or negative infinity) or if Mathcad indicates an error in the
rkfixed() function, either decrease the duration of the calcu-

x,v

F(t)

1
1

M x + M F


0
0

F(t)

K
g

MATLABs RungeKutta Solver ode45()

461

Fig. A7.6 Response of the state


variable v1g and FK to a step input
10N

2.938

15

11.085

2
10
v.1g

F.K

1
5
0

0.319

10

Fig. A7.7 Response of output


variables v1g, FK, FM and Fb to a
step input of 10N

2.938

20
t.MbK

25

15

10

11.085

F .K
v.1g

F .M
F .b

0.319

0
0

lation, or increase the number of calculations, to decrease the


time step. If the calculation is still numerically unstable,
check for a sign error in the Mathcad code or the reduction of
the state equations. A single sign error is sufficient to create
a numerical instability.
We will increase to duration of the calculation to 25sec
and the number of calculations to 2,000 for the rkfixed()
solver function.
ZMbK : rkfixed(x, 0, 25, 2000, D)
Extract the time and state variable vectors (in the computer
science meaning) from the array ZMbK.
t MbK : = ZMbK 0

v1g : = ZMbK 1

FK : = ZMbK 2

The output variables are calculated using values of the state


variables from the solution of the state equations. Often, the
state variables are included as output variables, as below.
0
v1g
1
0
F
0 1 v 0
K:=
1g + F
F
b
1

FK 1
F
b 0
0
b

10

20
t.MbK

2.59

25

MATLABs RungeKutta Solver ode45()


MATLABs general purpose RungeKutta solver function
is named ode45(). If it helps you remember the name, the
45 stands for fourth-order overall error, fifth-order local
error.
One of the arguments of the function ode45() is a userwritten MATLAB function which contains the state equations to be solved. A MATLAB user-written function resembles a short program or script except for the syntax of
the first line of the code is an instruction, which declares
the function. The declaration, or first line, of a MATLAB
user-defined function begins with the key word function
followed by an assignment statement where the left side is
a variable name and the right side is the function name followed by the functions argument list. The three dots or periods are ellipses, the mathematical symbol for continuation of a series. Ellipses are MATLAB's operator to continue
a statement on the following line.
function variableName = functionName(argument1,
argument2, argument3, additionalArguments)
The above syntax is generic for any function one wishes
to create. The function containing the state equations to be
solved by MATLABs RungeKutta solver ode45() expects

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

462

just two arguments. The first argument is the independent


variable t for time in our application. The second argument
is the variable name of the state vector x. Attempting to pass
other arguments to the function, which one can do with a
normal user-written MATLAB function, will not work with
the ode45() solving function. Parameter values and input
functions must be written into to the user-written function.
They cannot be passed to it as arguments.

Step Response of a Linear Mass-Damper System


We will illustrate creation of a user-written function and its
solution, first with a mass-damper and then with a springmass-damper system. The state equation of the mass-damper
system shown in Fig.3.13 is repeated below for reference:
dv1g
F (t ) dv1g b
b
F (t )
=
+ v1g
= v1g +
M
dt
M
dt
M
M
The state variable is v1g, the velocity of the mass.
The first step is to create a MATLAB user-defined function containing the state equations in MATLABs editor. The
function declaration begins with the keyword function. The
statement needs a variable name for the output of the function
and a name for the function itself. We will name the variable
dx and the function StateEq. There are two arguments of the
function. They must be the independent variablet for time in
our applications, and the state vector x in that order.
function dx = StateEq(t,x,) % Function Declaration
The first line of code within the function is to define the variable dx as a vector variable with the number of elements
equal to the number of state variables, one in this case. The
instruction used to create (or allocate) a vector or array and
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
one column. The instruction to create the column vector dx
with one element with the initial value of zero is:
% Create and zero the column vector dx with one row
dx = zeros(1,1);
The state equation is assigned to the element of the only element 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
x, x(1).
M = 3; % Mass
b = 2; % damping
F = 10; % Force
dx(1) = (-b/m)*x(1) + (1/m)*F; % State Eq.

An end statement completes the function. Collecting the


lines of code without the interspersed text:
function dx = StateEq(t,x,) % Function Declaration
% Create and zero the column vector dx with one row
dx = zeros(1,1);
M = 3; % Mass
b = 2; % damping
F = 10: % Force
dx(1) = (-b/m)*x(1) + (1/m)*F; %State Eq
end
The user-written MATLAB function is then saved using the
functions name and followed by the extension .m. In the
command window, the following command runs MATLABs
ode45() RungeKutta solver and writes the time data to the
vector T and the values of the state variable to vector X.
[T,X]=ode45(@functionName,[tstart tend],[vector
of the initial values of state variables])
The square brackets vectors. The elements in a vector can be
separated by a blank space or a comma. Rows are separated
by a semicolon. Note that the name of user- written function
with the state equations is preceded by the at symbol, @.
This is essential.
We will solve the state equation of the mass-damper system out to 12sec and use zero as the initial value of the
velocity of the mass.
[T,X] = ode45(@StateEq,[0 12],[0])
The vector T contains the times of the calculations and X
contains the values of the state variable at those times. The
vectors T and X are plotted with the command:
plot(T,X)
The result is shown in Fig.A7.8. Right clicking after having
selected an object brings up a context menu for editing the
plot. Select axes to label them and change their font size.
Increase the thickness of the output trace.

Step Response of a Non-Linear Mass-Damper


System
We can edit the state equations and add some types of nonlinearities, such as a power law viscosity for the damping.
We will square the velocity of the mass, x(1), in the state
equation.
function dx = NLStateEq(t,x,) % Function Declaration
% Create and zero the column vector dx with one row
dx = zeros(1,1);

MATLABs RungeKutta Solver ode45()


5

463

MassDamper Step Response

Linear and NonLinear MassDamper Step Responses

4.5

v1g, m/sec

V1g, m/sec

4
3.5
3
2.5
2
1.5
1

Nonlinear

3
2
1

0.5
0 0

Linear

t,sec

10

0
0

12

t, sec

10

12

Fig. A7. 8 Step response of a linear mass-damper system, M=3, b=2,


F=10

Fig. A7.9 Linear and non-linear step response of a mass-damper system, M=3, b=2, F=10

M = 3; % Mass
b = 2; % damping
F = 10: % Force
dx(1) = (-b/m)*x(1)^2 + (1/m)*F; %State Eq
end

if(t<2)
F = 5;
elseif(t<4)
F = -5;
else
F = 0;
end;
dx(1) = (-b/m)*x(1)^2 + (1/m)*F; %State Eq.
end;

Save this function as NLStateEq.m and run the RungeKutta


solver ode45() with different names for the output time and
state variable vector than used previously.
[Tnl,Xnl] = ode45(@NLStateEq,[0 12],[0])
We can now plot the linear and the non-linear step response
of the mass-damper system on the same plot with the command:
plot(T,X,Tnl,Xnl)

Save the edited function as PulseStateEq.m. Run the Runge


Kutta solver ode45() and plot the result with the commands:
[T,X] = ode45(@PulseStateEq,[0 12],[0])
plot(T,X)

Response of a Linear Mass-Damper System


Subjected to a Pulse Train

Response of a Mass-Damper System to Sinusoidal


Inputs

The independent variable t can be used in the state equations


for the input function. We will use an if, elseif, else structure
to create the input pulse:

We use the independent variable t to drive the system with


the input F (t ) = 10 sin(2t )

F (t ) = 5 us (t ) 10 us (t 2) + 5 us (t 4) .
Edit the first-order linear function StateEq(t, x) to add the
if, elseif, and else, shown below to calculate the input force
F(t).
function dx = PulseStateEq(t,x,) % Function Declaration
% Create and zero the column vector dx with one row
dx = zeros(1,1);
M = 3; % Mass
b = 2; % damping
% Force Pulse if, elseif, else

function dx = SinStateEq(t,x,) % Function Declaration


% Create and zero the column vector dx with one row
dx = zeros(1,1);
M = 3; % Mass
b = 2; % damping
dx(1) = (-b/m)*x(1)^2 + (1/m)*10*sin(2*t); %State Eq
end;
Save the function as SinStateEq.m, solve it with ode45() and
plot it.
[T,X] = ode45(@SinStateEq,[0 12],[0])
plot(T,X)

7 Vector-Matrix Algebra and the State-Space Representation of Dynamic Systems

464

Input F(t)=10sin(2t) Linear MassDamper System


ode45( ) option refine set to 8

Pulse Response of Linear MassDamper System


2

3
2

v1g, m/sec

v1g, m/sec

1.5

0.5
0
0.5

0
1

1
1.5 0

10

12

t, sec
Fig. A7.10 Response of the linear mass-damper system, system,
M=3, b = 2, to the input pulse F(t)

if(t<2)
F = 5;
elseif(t<4)
F = -5;
else
F = 0;
end;

t, sec

10

12

Fig. A7.12 Response of the linear mass-damper system, M = 3, b = 2,


to the input F(t)=sin (2t). The refine option was increased from its
default of four to eight

than decreasing the time step. The refine option instructs


the ode solver interpolate within a time step. To the user, it
appears that the time step has been decreased. The default
valve of refine for ode45() is four. We will double the value
to eight. The syntax is:

Input sin(2t), Linear MassDamper System


2.5

options = odeset('refine',8)
The variable name, options, is the last argument in the argument list of ode45

2
1.5

v1g, m/sec

[T,X] = ode45(@SinStateEq,[0 12],[0],options)

0.5
0

The refined output is plotted in Fig.A7.12. The appearance


of the trace is improved. It no longer has straight segments.

0.5
1
1.5
2

t, sec

10

12

Fig. A7.11 Response of the linear mass-damper system, M=3, b=2,


to the input F(t)=sin(2t). Notice the slightly blocky shape of the trace

The quality of the response, Fig.A7.11, is disappointing.


What should be pleasingly smooth sine curves are not. There
are straight segments.
We must delve into the options available for ode45() reduce the blockiness. MATLAB has a built in function named
odeset() for setting options for their ordinary differential
solvers. The function odeset() exists because there are 22 options for ode solvers and you cannot pass just a single option
value to the solver. It expects all 22 arguments. The function odeset() will accept a single option and its value. It then
creates a file will the remaining option at their default values.
The set of options is assigned to a variable name.
The option we will change is titled refine. MATLAB
has a faster routine for refining the result of an ode solver

Step Response of a Spring-Mass-Damper System


Our final example will be to solve for the step response of
the state variable and a number of output variables of the
second-order linear spring-mass-damper system, shown in
Fig.7.5.
We will illustrate its use with the state equations of the
mass-spring-damper system:
b
d v1g
=
M
dt FK
K

1
1
v1g
M + M F
F

0 K 0

The MATLAB code requires individual equations. Expanding the matrix form:
dv1g

1
1
b
v1g
FK +
F
M
M
M

dt
dFK
= Kv1g
dt

References and Recommended Reading


20

465

SpringMassDamper Step Response

Force FM and Fb, Step Response SpringMassDamper System


10

FK

FM
5

FM, N
Fb, N

V1g, ms/sec
FK, N

15
10
5

V1g

Fb

0
5
0

10
10

20

30

40

50

t, sec

60

70

80

90

100

Fig. A7.13 Step response of the linear spring-mass-damper system,


M= 3, b = 0.2, K=1 to the input F = 10

It is always worth the time to define which state variable is


which. In this case, v1g x1 and FK x2 .
We will name the variable dx and the function StateEqs.
The two arguments of the function must be t for time and x
for the state vector, in that order.
The first line of code within the function is to define the
variable dx as a vector with the number of elements equation
to the number of state variables, two in this case. The instruction to create the column vector dx with two elements with
the initial values of zero is:
function dx = StateEqs(t,x,) % Function Declaration
% Create and zero the column vector dx with one row
dx = zeros(1,1);
M = 3; % Mass
b = 0.2; % damping
K = 1; % Spring
F = 10; % Force
dx(1) = (-b/M)*x(1) + (-1/M)*x(2) + (1/M)*F;
dx(2) = K*x(1);
end
The output equations are purely algebraic equations. They
are solved after the state equations have been solved and the
state variables are known. We must expand the output equations from vector-matrix form into a set of equations in order
to program them.
0
v1g 1
0
F 0 1 v

K=
1g + 0 F
FM b 1 FK 1


0
Fb b 0

v1g = v1g
FK = FK
FM = bv1g FK F
Fb = bv1g

20

40

60

t, sec

80

100

Fig. A7.14 Step response of the output variables FM and Fb of a linear


spring-mass-damper system, M=3, b=0.2, K = 1 to the input F=10

The time values are in the vector variable T. Its length can be
determined by the function length(). The state variables v1g
and FK are the columns in the state vector X and should be
extracted and assigned to individual vector variables. MATLABs colon operator,:, is a wildcard to force a variable to
take on every value in a range. The colon operator successively takes on the values of the row indices, allowing the
columns of the array to be copied to vectors without the use
of a for end loop.
v1g = X(:,1); % Extracts column 1 from the array X
FK = X(:,2); % Extracts column 2 from the array X
The output variables are calculated with a for loop.
b = 0.2; % Damping
F = 10; % Step input
N = length(T); % Determines length of time vector T
FM = zeros(N); % Creates and zeros vector FM
Fb = zeros(N); % Creates and zeros vector Fb
% for loop to calculate output equations
for n=1:N;
FM(n)=-b*v1g(n)-FK(n)+F;
Fb(n)=b*v1g(n);
end

References and Recommended Reading


Ogata, K (2009) Modern control engineering, 5th edn. Prentice-Hall,
Englewood Cliffs
Rowell D, Wormley DN (1997) System dynamics: an introduction.
Prentice-Hall, Upper Saddle River

Finite Difference Methods and MATLAB

Abstract

Finite difference approximations are the foundation of computer-based numerical solutions


of differential equations. Numerical solutions, also known as numerical methods, are essential 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 RungeKutta algorithm is
also more involved. Fortunately, the use of the RungeKutta method is straightforward. The
Appendix to the chapter is a brief introduction to computer programming and programming
in MATLAB for those who need it.

8.1Finite Difference Approximation


of Differential Equations
We will approximate the differential equations, which describe the dynamics of energetic systems using a finite difference approximation. Recall from your introduction to
calculus, the definition of a derivative of a variable, say v,
with respect to time as
lim

t 0

v dv
=
t dt

In a finite difference approximation of the derivative, the infinitesimal t produced by evaluating lim is replaced with
t 0
a small but finite t:


v dv

t dt

It is good practice to test the effect of the size of the time


step on the results of the calculation, by reducing t by a
significant fraction, and rerunning at least part of the calculation. If there is no difference, then the time step is small
enough. If the two calculations differ, reduce the time until
there is no significant difference in the results.
A finite difference approximation of a differential equation is created by replacing all of the differentials with finite differences. For example, the differential equation with
force, F(t), is the input and velocity, v(t), is the output
F (t ) = 3
which is approximated as

(8.1)

How small a t is small enough for a reasonable approximation of the derivative? That depends on the time scale of the
dynamic response of the system. Using the time constant,
, of a first-order system and the period, T, of an oscillatory
second-order system as time scale of the respective systems
dynamic responses, the upper limits of t should be approximately t / 20 and t T / 200 so that plots that are supposed to be smooth curves are smooth curves.

dv(t )
+ v(t )
dt

F (t ) = 3

v(t )
+ v(t )
t

Notice this equation expresses time in two different forms.


The variables, F(t) and v(t), are written as functions of continuous time, but the derivative is approximated by using a
finite (or discrete) time, t. We cannot evaluate this equation
at any arbitrary time, t, but only at times which are multiples
of our finite time, t,


t = n t

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_8, Springer Science+Business Media New York 2014

(8.2)
467

8 Finite Difference Methods and MATLAB

468
Fig. 8.1 Forward-stepping finite
difference calculation as a function of discrete time

v((n+1)t)

v((n+1)t)

v(nt)

Velocity
Continuous
Variable

v(nt)

v(nt)
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
0

(n-2)t

t
(n-1)t

t
nt

(n+1)t

Time
Discrete Variable
where n is an integer. Time calculated as a multiple of a time
step, t, is known as discrete time, as opposed to continuous time. Substituting n t for t:


F (n t ) = 3

v(n t )
+ v(n t )
t

(8.3)

This approximation of the differential equation, expressed


using finite differences, is called a difference equation.

8.2Euler Method, Forward Stepping,


Finite Difference Algorithm
A recursive calculation is an algorithm, in which the new
value is calculated from past values. A forward-stepping, finite difference calculation is a recursive algorithm, where the
new value calculated from the past values is the value of the
output variable for the next forward step in time.
We can rearrange the difference equation, Eq.8.3, to
allow us to calculate the change in the output variable, v,
at time, t = n t


v(n t ) =

F (n t )
v(n t )
t
t
3
3

(8.4)

v(n t ) is the change in the velocity, v(n t ), over the finite


time, t, from time, t = n t, to time, t = (n + 1) t. Hence,
adding the current value of v to the forward change v
yields the next value of v:
v((n + 1) t ) = v(n t ) + v(n t )
(8.5)

Substituting our expression for v(n t ) , Eq.8.4, into Eq.8.5


yields


v((n + 1) t ) =

F (n t )
v(n t )
t
t + v(n t ) (8.6)
3
3

The calculation steps forward in time, by calculating the


change in velocity, v(n t ), over the time period, t, and
then adding it to the velocity at the beginning of the time
period, to calculate the velocity at the end of the time period.
The time period, t, is known as the time step. The forward stepping solution of a difference equation is illustrated
in Fig.8.1.
The time axis is identified as a discrete variable, which
means it can only take on certain values. In this example,
time must be an integer product, or multiple, of the time step,
t. In contrast, the velocity axis is identified as a continuous variable. There is no restriction on the value the velocity variable can take.
The forward-stepping Euler algorithm will be illustrated
by calculating two steps of the difference equation

v(n t ) =

F (n t )
v(n t )
t
t
3
3

To begin the calculation, we must know the initial value


of the output variable. We will assume v(0+ ) = 0. We will
impose a step input on the system of a magnitude of ten,
F (t ) = 10. Lastly, we will use a time step, t = 0.5.
The first calculation is the change of the output variable
at time, t=0, v(0.0):

v(0.0) =

F (0.0)
v(0.0)
t
t
3
3

10
0
0.5 0.5 = 1.67
3
3

v(0.0) =

8.2 Euler Method, Forward Stepping, Finite Difference Algorithm

469

Add the change of the output value, v(0.0), to its existing


value, v(0.0), to yield the value of the output variable for
next discrete time, t + t = 0.0 + 0.5 = 0.5. This is the forward step in time to the velocity at time, v(0.5):

v(n+1)

Velocity

v(n)

v(n)

v(0.5) = v(0.0) + v(0.0)

v(n-1)

v(0.5) = 0 + 1.67 = 1.67

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) :
F (0.5)
v(0.5)
v(1.0) =
t
t
3
3
10
1.67
0.5 = 1.39
v(1.0) = 0.5
3
3
Add v(0.5) and v(0.5) to again step forward in time to
the velocity at time, v(1.0):
v(1.0) = v(0.5) + v(0.5)
v(1.0) = 1.67 + 1.39 = 3.06
Recursive algorithms must retain previous values of the calculation in order to calculate the next value. In our applications, i.e., the solution of a system equation or a set of state
equations, we will have to create a complete record of the
results of the calculation in order to plot them. We will store
the values of the output variable and time in vectors. This
raises the question how to modify the calculation we just
performed to store the values we need in vectors. Although
we wrote the finite difference equation, Eq.8.4, in discrete
time, t=nt, we do use the value of the discrete time in the
calculation. It is sufficient to merely identify the previous
value of the output variable. A step in the calculation can be
identified by either the time corresponding to the step, nt,
or, more simply, by just value of the integer variable, n. It is
easier to program the calculation as a sequence with the step
number, n, as the index of the vector variables.
The standard mathematical notation for an element in a
series is a subscript, as is used in Mathcad to index vector
elements. Unfortunately, computer code uses neither subscripts nor superscripts. The characters of computer code are
those found on a standard keyboard. The syntax used to denote a vector index in MATLAB is to append the subscript in
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
vector is identical to the notation used to denote a function
of one variable.
For example, consider the same first order written as a
function of discrete time:

v(n t ) =

F (n t )
v(n t )
t
t
3
3

v(n+1)

v(n)
v(n-1)
v(n-1)
v(n-2)
v(n-2)

n-2

{
n-1

Step Number

n+1

Fig. 8.2 Forward-stepping finite difference calculation, referencing a


value by its step number, n

and written using the counter variable, n, as the index for the
input vector, F, and the output vector, v:


v(n ) =

F (n )
v(n )
t
t
3
3

(8.7)

The equation written as a function of discrete time is how a


human may think of a step in the calculation, but the calculation is programmed using the vector form of Eq.8.7, shown
in Fig.8.2. Discrete time is not needed to perform the calculation, only to plot the result on a time axis.
Since computer code is restricted to standard keyboard
characters, there are no Greek characters. We will rewrite
the finite difference approximation of the derivative using
D for in x(n). Likewise, we will use T rather than t.
Be careful with the variable, T. T can also represent the
period of an oscillation. Comments in the MATLAB code
which identify variables are a great help to understanding an
algorithm.

8.2.1MATLAB Programming of the Euler


Method, First-Order System
The first MATLAB program, or script, will be the solution of
the finite difference we just solved manually:
F (t ) = 3

dv(t )
+ v(t )
dt

To make the program of more general use, we will code it in


the standard notation of a state equation, using u(t) for the
input; x(t) for the state variable; a for the coefficient multiplying the state variable; and b for the coefficient multiplying the input:
bu (t ) =

dx(t )
+ ax(t )
dt

8 Finite Difference Methods and MATLAB

470

Rearranging both equations into the standard form of a state


equation

Start

dv(t )
1
1
dx(t )
= v(t ) + F (t )
= ax(t ) + bu (t )
dt
3
3
dt

Input values
N, t end , a, b, x(1), u

yields a = 1/3 and b = 1/3.


Calculate time step
t end
T = ___
N

We also need to specify the number of iterations, or steps N,


that the program is to execute and the time step, T. Rather
than assign a value to the time step, T, it is more convenient
to specify the duration, tend, of the calculation and let the
computer calculate the time step, T:


t T =

tend
N steps

Initialize index variable n


and time vector t(n)
n=1
t(1) = 0

(8.8)

In order to plot the results, we must create vectors of time


and the output, which, in this case, is the state variable, x.
The vectors must have the same number of elements. MATLABs 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.


t (n) = (n 1) T

Dx = a x(n) T + b u T

Calculate forward step


x(n+1) = Dx + x(n)

(8.9)

Equation8.9 illustrates an unfortunate consequence of not


being able to use subscripts in computer code. The parenthetical n in t(n) is the index of the vector, t. The parenthetical (n1) 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
with the number of steps, N, and the execution of the loop is
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,
as shown in the following code.
% 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
b = 1/3 % Coefficient of input variable
x(1) = 0 % Initial value of the state variable
u = 10 % Input function, a step of 10
T = t_end/N % Time step
t(1) = 0 % Initial value of the time vector
for n = 1:N; % Finite difference loop
Dx = a*x(n)*T+b*u*T;
x(n+1) = Dx+x(n);
t(n+1) = n*T;
end
plot(t,x)

Calculate the change Dx

Calculate time t(n+1)


t(n+1) = n T

Increment counter by one


n=n+1

No

n=N?

Yes
Plot results

Finish

Fig. 8.3 Flowchart of the Euler method solution of a first-order systems state equation

Note that the change in the state variable is now a scalar


variable, Dx, rather than a vector variable, Dx(n). There is
no need to store the change in the state variable for each
time step, unless there is a desire to plot it. The scalar variable, Dx, is assigned the current value for each iteration of
the loop.
An alternative method uses a while loop to control the
number of iterations.

8.2 Euler Method, Forward Stepping, Finite Difference Algorithm


Fig. 8.4a Force source acting on a spring-mass-damper
system. b Linear graph of the
spring-mass-damper system

471

x,v

F(t)

b
M

F(t)

K
g

Frictionless rollers
% Euler Method, First-order system, while loop
N = 2000 % Number of steps
t_end = 20 % End time of calculation
a = -1/3 % Coefficient of state variable
b = 1/3 % Coefficient of input variable
x(1) = 0 % Initial value of the state variable
u = 10 % Input function, a step of 10
T = t_end/N % Time step
t(1) = 0 % Initial value of the time vector
while (N+1) > n;
Dx= a*x(n) T+b*u*T;
x(n+1) = Dx+x(n);
t(n+1) = n*T;
n = n+1;
end
plot(t,x)

The state equations expressed in vector-matrix form are:

8.2.2Euler Method Solution of Second-Order


State Equations

Multiplying both sides by the time step, t, yields a set of


simultaneous difference equations:

We will illustrate the solution of a set of state equations using


the Euler method for the force source spring mass-damper
system, Fig.8.4.
Energetic Equations
Continuity F FM FK Fb = 0
Compatibility v1g = v1g
Elements FM = M

dv1g
dt

Energy E sys = E M + E K

dFK
= Kv1g Fb = bv1g
dt
F2
1
E M = Mv12g E K = K
2
2K

The state variables in this formulation are the energy storage


variables, the velocity of the mass, v1g, and the force acting
through spring, FK. The reduction of the energetic equations
yields the following state equations:
dv1g

1
1
b
= v1g
FK +
F
dt
M
M
M
dFK
= Kv1g
dt

b
d v1g
=
M
dt FK
K

1
1
v1g

+
M M F
F

0 K 0

We can derive the finite difference equations for this system, by either operating on the individual state equations,
or on the state equations in vector-matrix form. However,
we need the individual equations to program the equations.
The first step is to approximate the derivatives as finite differences:

v1g
1
1
b
= v1g
FK +
F
M
M
M
t
FK
= Kv1g
t

b
1
1
v1g t
FK t +
F t
M
M
M
FK = Kv1g t

v1g =

As we did with the Euler solution for a single-state equation, we will write the program for a system of two state
equations in standard notation. We will begin in vectormatrix form.
d x1 a11
=
dt x2 a21

a12 x1 b1
u
+
a22 x2 b2

Our first step is to approximate the differential operator


with the finite difference operator. In other words, replace
d with .

x1 a11 a12 x1 b1
u
=
+
t x2 a21 a22 x2 b2

8 Finite Difference Methods and MATLAB

472

Next, multiply both sides by t and distribute the finite difference 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.
x1 a11
x = a
2 21

Start

Input values
N, t end , a11, a12, a21, a22
b1, b2, x1(1), x2(1), u(n)

a12 x1
b
t + 1 ut

a22 x2
b2

Calculate time step


t end
T= ___
N

Expand the vector-matrix notation into a set of two simultaneous finite difference equations, in order to write the MATLAB code.


x1 = ( a11 x1 + a12 x2 + b1u ) t


x2 = ( a21 x1 + a22 x2 + b2 u ) t

Change the notation and express Eqs.8.10 with the characters available on a standard keyboard, without or subscripts. Rename the variables, x1 x1, x1 Dx1, x2 x 2,
and x2 Dx 2. Rewrite the coefficients without subscripts, as a11 a11 and b1 b1 Use the vector index, n,
and the parenthetical notation for a vector subscript used
in MATLAB to rename the variables and coefficients of
Eqs.8.10:


Dx1 = ( a11 x1(n) + a12 x 2(n) + b1 u (n) ) T

Dx 2 = ( a 21 x1(n) + a 22 x 2(n) + b 2 u (n) ) T

Initialize index variable n


and time vector t(n)

(8.10)

n=1

Calculate Dx1 and Dx2


Dx1 = (a11 x1(n) + a12 x2(n) + b1 u(n)) T
Dx2 = (a21 x1(n) + a22 x2(n) + b2 u(n)) T

Calculate forward step


x1(n+1) = Dx1 + x1(n)
x2(n+1) = Dx2 + x2(n)

(8.11)

Equations8.11 are both evaluated before the vectors, x1(n)


and x2(n), are updated by adding the change, Dx1 and
Dx2, to the current values, x1(n) and x2(n), to yield the value
for the next, n+1, calculation.


x1(n + 1) = Dx1 + x1(n)


x 2(n + 1) = Dx 2 + x 2(n)

(8.12)

The index, n, is then incremented and the loop repeated, until


n=N. Combine Eqs.8.11 and 8.12 as:
 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)

t(1)=0

Calculate time t(n+1)


t(n+1) = nT

Increment counter
n=n+1

No

n=N?
Yes
Plot results

Finish

(8.13)
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
parameters of the spring-mass-damper system.
%Euler Method, Second-order system,
%spring-mass-damper
N = 2000 % Number of steps

Fig. 8.5 Flowchart of the Euler method algorithm solution for a twostate system

t_end = 100.0 % Duration of calculation


M = 3.0 % Mass
K = 1.0 % Spring constant
B = 0.2 % Damping coefficient
% Coefficient matrix elements
a11 = -B/M
a12 = -1/M

8.2 Euler Method, Forward Stepping, Finite Difference Algorithm

force, when v1g = 2 m/sec. We will model a thickness of the


lubricating film as constant and independent of the velocity
of the mass. Assume a thickness of h = 0.02 mm for the lubricating film. We need an area for the shearing surfaces
and will use A = 10 cm 10 cm = 0.01 m 2 . We begin by
equating the damping forces at v1g = 2 m/sec to calculate
the value of 0:

100
80

()

473

60

Viscosity 40
20

10

15

Shear Strain Rate,

20

v1g 2
Fb = bv1g = A 0 v1g
h

Fig. 8.6 Power law viscosity, Eq.8.14

0 =
a21 = K
a22 = 0.0
% Input vector Elements
b1 = 1/M
b2 = 0.0
%Initial values of the state variables
x1(1) = 0.0 % Velocity v1g
x2(1) = 0.0 % Force FK
u = 10.0 % Input step magnitude
T = t_end/N % Time step
t(1) = 0.0 % Initial value of time vector
for n = 1:N;
Dx1 = (a11*x1(n) + a12*x2(n) + b1*u)*T;
Dx2 = (a21*x1(n) + a22*x2(n) + b2*u)*T;
x1(n+1)= Dx1 + x1(n);
x2(n+1)= Dx2 + x2(n);
t(n+1) = n*T;
end;
plot(t,x1,t,x2)

N sec

0.2
2 m 2
m
sec
=
2

0.01 m
0.00002 m
1

Rearrange the elemental equation for the power law damper


to collect the velocity term
3

Fb = bv1g

v1g 2
Fb = 0 A v1g
h

52
A 0 5
v1g
Fb = 0 A 3 = 3 v1g 2

h 2 h 2
yielding
A 5
0
Fb = 3 v1g 2
2
h

Fb = bN .L.v1g 2

where

b=

A 0
3

h2

The state equations of the spring-mass-damper system are:

MATLAB and Mathcad both include a built-in Runge


Kutta solver function. The reason to program a finite difference solver is for the inclusion of non-linear system elements. In this example, rather than assuming that lubricating
film has ideally linear viscous friction, we will use a power
law model for the viscosity of the film, Fig.8.6 and Eq. 8.14:
3

( ) = 0 2

3
2

0 = 6.3 10 7 kg m 2 sec 2

8.2.3Example: Euler Method Solver,


Non-Linear State Equation

b v1g
A h

v1g 2
b = A 0
h

v1g
1
1
b
= v1g
FK +
F
M
M
M
t
FK
= Kv1g
t
3

Substitute 0 / h 2 for b and raise v1g to the 5/2 power:

(8.14)

v1g
1
=
M
t

We will derive a power law viscosity, such that the damping force for the linear damper with damping coefficient
b = 0.2 N sec/m and the power law damper have the same

FK
= Kv1g
t

A 5 1
1
3 0 v1g2
FK +
F
M
M
2
h

8 Finite Difference Methods and MATLAB

474

we can rewrite the expression into the product of x and the


absolute value of x raised to a fractional power.

Multiply both sides by the time step, t

1
M

v1g =

A 5 1
1
3 0 v1g2
FK +
F t
M
M
2

x = x x

FK = Kv1g t
Our second-order state equation solver is written in terms of
coefficients, ar,c, and state variables, x1 and x2. Hence,
1 A
a11 = 3 0
M 2
h

x2 = ( a21 x1 + a22 x2 + b2 u ) t

(8.15)

If we were to code this set of finite difference equations, and


run it in MATLAB, we would receive an error message
Warning Imaginary parts of complex X and/or Y
arguments ignored.

The problem is that the velocity of the mass, v1g , changes


sign, as it oscillates about zero. The power law viscosity expression yields a real value when the velocity is positive, but
a complex number when the velocity is negative. Raising a
number to an odd power does not change its sign. Consequently, when the velocity is negative, we are attempting to
take the square root of a negative number.
The power law model viscosity model, Eq.8.14, should
have been written using the absolute value of strain rate to
avoid computational error:

( ) = 0

3
2

The state equation was rewritten to calculate the fluid strain


rate in terms of the mass velocity, v1g . We must maintain the
sign of the damping force. We cannot take the absolute value
of velocity in the damping force expression, in this case:
A
Fb 3 0 v1g
2
h

5
2

Using the relationships:

( )

x = x

and

x ( +1) = x x

This allows us to maintain the sign of velocity and take the


square root:



5


x1 = a11 x1 2 + a12 x2 + b1u t

=x x

A
Fb = 3 0 v1g v1g
2
h

3
2

(8.16)

The state equations are now:

and


x1 = a11 x1 x1

3

2

+ a12 x2 + b1u t

(8.17)

x2 = a 21 x1 + a 22 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.
In this case, we take the absolute value of an element of a
vector.
MATLAB code for the Euler method solver for a secondorder system with the power law viscosity follows:
% Second-order system with power law viscosity
% Euler Method state equation solver
N = 2000 % number of steps
t_end = 40.0 % duration of simulation
% Parameter values
M = 3.0 % Mass
K = 1.0 % Spring constant
mu = 6.3*10^-9 % Viscosity
h = 0.02*10^-3 % Height
% Coefficient matrix elements
a11 = (-1/M)*(mu/h^(3/2))
a12 = -1/M
a21 = K
a22 = 0.0
% Input vector elements
b1 = 1/M
b2 = 0.0
% Initial values of state variables
x1(1) = 0.0 % mass velocity v1g
x2(1) = 0.0 % Force acting through spring
%
u = 10.0 % step input F
T = t_end/N % time step
t(1) = 0.0 % initial value of time vector
% Difference equation loop
for n = 1:N;
dx1 = (a11*x1(n) (abs(x1(n))^(3/2)) + a12*x2(n)
+ b1*u)*T;

8.3 User-Written MATLAB Functions

Kinetic Friction

475

Friction
Force
FC
Static Friction
Velocity

Fig. 8.7 Static and kinetic Coulomb friction model. FC is the maximum value of static Coulomb friction

dx2 = (a21*x1(n) + a22*x2(n) + b2*u)*T;


x1(n+1)= dx1 + x1(n);
x2(n+1)= dx2 + x2(n);
t(n+1) = n*T;
end;
plot(t,x1,t,x2)

8.3 User-Written MATLAB Functions


8.3.1 Static-Kinetic Coulomb Friction Model
Coulomb friction is a more strongly non-linear phenomenon
than power law viscosity, and significantly more difficult to
incorporate into a dynamic model. We will use the familiar,
simplified model of static Coulomb friction and kinetic Coulomb friction, shown in Fig.8.7.
Static and kinetic Coulomb friction both manifest as functions of the normal force, acting across two surfaces in contact. Coulomb friction results from asperities (microscopic
high spots) on one surface making contact with asperities
on the other surface. Coulomb friction opposes shear motion
between solid surfaces. Static friction opposes impending
motion. Kinetic friction opposes actual motion.
In the abstract, it seems reasonable to define static friction as applicable for v(n) = 0, and kinetic friction as applicable when v(n) 0. However, this is not practical. We
should not use the criterion that the velocity of the mass
equals zero, as the condition to distinguish between kinetic and static frictions. An MATLAB if statement contains 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
the if statement are executed. If the condition is false, then
they are not. If we test whether two variables are equal,
and the variables are numerical variables (as opposed to
logical variables, which are Boolean), the two quantities
will be compared to the numerical precision inherent in

those variables. A mass velocity of v(n) = 0.001 mm/sec,


or a micron per second, is certainly small enough that
we would perceive it to be zero. However, the following
MATLAB code:
% Set value of v(n) to 1 micron/sec = 0.000001 m/sec
v(n) = 0.000001
% Set logical variables static and kinetic false
static = false
kinetic = false
% Test velocity and set friction condition true
if(v(n) ==0)
static = true
else
kinetic = true
end
yields kinetic=true, when the condition v(n)==0 is eval
uated since the velocity is not exactly, to the numerical
precision of the variable, zero. We eliminate the problem of
needing a value to be exactly zero by defining a velocity
limit, which is sufficiently small that it is, for practical purposes, zero, and then test whether the velocity in question is
above or below that limit. We must use the absolute value of
v(n) to prevent negative velocities with magnitudes greater
than the zero limit from satisfying the inequality.


v(n) < vZeroLimit

(8.18)

The following MATLAB code with a limit defined for zero


velocity yields static=true when the magnitude of the velocity is less than the zero limit.
% Define zero velocity limit to 0.1 mm/sec
vZeroLimit = 0.0001
% Set value of v(n) to 1 micron/sec or 0.000001 m/sec
v(n) = 0.000001
% Set logical variables static and kinetic false
static = false
kinetic = false
% Test velocity and set friction condition true
if(abs(v(n)) < vZeroLimit)
static = true
else
kinetic = true
end
Kinetic friction is modeled as constant force opposing motion. We determine the direction of the motion by the sign
of the velocity variable. All programming languages have
a function to determine the sign of a number, generically
known as the signum function, contraction of sign of
number. MATLABs signum function is sign(x). If the

476

sign of x is positive, the function returns one. If the sign


of x is negative, the function returns negative one. The
following MATLAB statement calculates the kinetic coulomb friction.
% if kinetic, calculate kinetic friction
if(kinetic)
Fkinetic = mu_k*F_N*sign(v(n))
end
Static Coulomb force is more difficult to incorporate in a dynamic model for two reasons. First, because static Coulomb
force opposes impending motion, the model must predict the
direction of the impending motion, in order to assign the correct direction to the friction force. Second, the magnitude of
static friction is variable. The magnitude indicated in the plot
as FC, as per Fig.8.7, is the maximum static friction, which
can be mobilized to resist impending motion. The static
friction mobilized falls in the range of FC.
The direction of impending motion is determined by the
sign of the sum of the forces acting on a mass. Suppose an
applied force F and a spring force FK act on a mass, in addition to the friction force. The sign of the direction of impending motion is:
% If static, determine sign of impending motion of mass
if(static)
impending = sign(F+FK)
end
There are two cases for calculating the magnitude of the
static friction. If the value of completely mobilized friction, FCMax = s FN , is less than the sum of the non-friction
forces acting on the mass, then FCMax is the friction force.
The second case applies, when the sum of the non-friction
forces acting on the mass is less than the completely mobilized friction, FCMax. The maximum permissible value of the
static Coulomb friction force is the force, which produces
static equilibrium, and, hence, equals the opposite of the
sum of the other forces acting on the mass. If the static
Coulomb friction calculated by the algorithm exceeds the
sum of the non-friction forces acting on the mass, then the
mass would accelerate under the action of friction, which
is nonsense.

8.3.2 Programming a Function in MATLAB


Although we could include the instructions we need to determine the Coulomb friction in our MATLAB program, we will
write a MATLAB function instead. The term, function, has
the same meaning in computer programming, as it does in
conventional mathematics. Mathematical functions, such as

8 Finite Difference Methods and MATLAB

the sine function, sin(x), are inputoutput relationships. The


input is the parenthetical quantity, known as the argument.
The function yields, or returns, a value. Software functions
are also inputoutput relationships. The input for a software
function, such as the absolute value function, abs(x), is also
the argument. The output of the function is the value it returns.
Computer languages allow users to write their own functions. A computer function differs little from a computer program. There are two advantages to using a function. First,
functions have fewer instructions and are easier to debug than
programs. Second, functions allow easy reuse of code in the
same program or a different program.
The syntax of using (or calling) a function in a computation follows mathematical convention. We will name
our Coulomb friction function, Coulomb. We will need to
Table 8.1 Function Coulomb arguments
Velocity of the mass
Sum of the input force and the spring force
Normal force acting across the surface
Static coefficient of friction
Kinetic coefficient of friction

v(n)
F + FK Fsum
FN

s
k

pass the function the following arguments:


Write functions with the same process and logic as a
MATLAB script (program). First, make a flow chart of the
logic, then encode it as MATLAB instructions. The flow
chart is Fig.8.8.
A function must begin with a MATLAB statement called
a declaration, which tells MATLAB that the following
code is (1) a function which will return, or provide, the
value of a given variable; (2) has a given function name;
and (3) uses the given list of variables as arguments in its
algorithm:
function return_variable
= function_name(argument or argument list)

In standard mathematical notation, we would calculate the


Coulomb friction using our function Coulomb, as
FC = Coulomb (v(n), Fsum , FN , s , k )
The declaration of our Coulomb friction function in MATLAB provides the same information, preceded by the keyword, function
function Fc=Coulomb(V, Fsum, Fn, mu_s, mu_k)

Note that function declaration expresses the velocity as V,


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


Coulomb friction model with
indeterminate static friction

Function Coulomb
v(n), Fsum ,FN , s , k

Define zero velocity


v zero = 0.001

Yes

|v(n)| > v zero

No
Calculate kinetic
Coulomb friction
FC = -k FN sign(v(n))

Calculate max limit of


static Coulomb friction
FC = s FN

FC >= |Fsum|

Yes

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
of the velocity of the mass to the function. We will not
pass the entire vector, v, to the function.
The following is the same logic, shown in the Fig.8.8
flowchart, coded using an else statement.
% Function Coulomb
% Returns static or kinetic friction
%
function Fc = Coulomb(V,Fsum,Fn,mu_s,mu_k);
vZeroLimit = 0.001;

%
% If the absolute value of the velocity is greater than
% the zero limit then calculate the kinetic Coulomb
% friction, else calculate the static Coulomb friction
%
if (abs(V) > vZeroLimit);
Fc = -mu_k*Fn*sign(V);
else;
%
% Calculate the static Coulomb Friction
%
Fc = mu_s*Fn;

8 Finite Difference Methods and MATLAB

478

if (Fc >= abs(Fsum));


Fc = -Fsum;
else;
Fc = -sign(Fsum)*Fc
end;
end;
Here is the same logic coded, using nested if statements.
% Function Coulomb_ifs
% Returns static or kinetic friction
%
function Fc = Coulomb_ifs(V,Fsum,Fn,mu_s,mu_k);
%
vZeroLimit = 0.001;
%
% If the absolute value of the velocity is greater than
% the limit then zero limit then calculate the
% kinetic Coulomb friction.
%
if (abs(V) > vZeroLimit);
Fc = -mu_k*Fn*sign(V);
end;
%
% If the absolute value of the velocity is less than or
% equal to the zero limit then calculate the
% static Coulomb friction
%
if (V <= vZeroLimit);
Fc = mu_s*Fn;
if (Fc >= abs(Fsum));
Fc = -Fsum;
end;
if (Fc < abs(Fsum));
Fc = -sign(Fsum)*Fc
end;
end;
The function is used, or called, in our state equation solver
program in the same manner, that functions are used in conventional mathematical notation. Return to the state equations
of the spring-mass-damper system expressed in terms of finite
differences:

v1g
1
1
b
= v1g
FK +
F
M
M
M
t
FK
= Kv1g
t
We see that the Coulomb friction force, FC, can be included in a state equation for the velocity of the mass, either as
an additional term with the coefficient, 1/M, or summed
with either the spring force, FK, or the input force, F. The

other choice is whether to use an assignment statement to


define a Coulomb friction variable, say Fc, or to include
the function call in the state equation. The following code
uses the latter method. The function Coulomb is called
and the value returned is assigned to the variable, Fc.
The friction force is then summed with the input force,
F. First, express the logic in conventional mathematical
notation:
Fc = Coulomb(V , Fsum, Fn, mu _ s, mu _ k )
b
1
1

v1g
FK + ( Fc + F ) t

M
M
M
FK = Kv1g t

v1g =

Now, change notation to express these equations using the


characters available on a standard keyboard. Express the
variables, x1, x1 , x2, and x2 , as vectors with the index, n,
using the standard notation for state equations.
% Second-order system Euler method state equation
% solver with linear damping and static and kinetic
% friction
%
N = 2000 % number of steps
t_end = 100.0 % duration of simulation
%
% System parameter values
M = 3.0 % mass
K = 1.0 % spring constant
B = 0.2 % damping coefficient
mu_s = 0.03 % coefficient static of friction
mu_k = 0.02 % coefficient kinetic of friction
%
Fn = M * 9.81 % Normal force
%
% Coefficient matrix elements
a11 = -B/M
a12 = -1/M
a21 = K
a22 = 0.0
% Input matrix elements
b1 = 1/M
b2 = 0.0
% Initial values of state variables
%
x1(1) = 0.0 % mass velocity v1g
x2(1) = 0.0 % Force acting through spring FK
%
u = 10.0 % step input F
T = t_end/N % time step
t(1) = 0.0 % initial element of time vector
%
% Difference equation loop

8.4 RungeKutta Method

479

for n = 1:N;
V = x1(n); %Current mass velocity
Fsum = x2(n) + u; %Net force w/o Coulomb friction
% Function call
Fc = Coulomb(V,Fsum,Fn,mu_s,mu_k);
% Euler Algorithm
Dx1 = (a11*x1(n) + a12*x2(n) + b1*(Fc + u))*T;
Dx2 = (a21*x1(n) + a22*x2(n) + b2*u)*T;
x1(n+1)= Dx1 + x1(n);
x2(n+1)= Dx2 + x2(n);
t(n+1) = n*T;
end;
plot(t,x1,t,x2)

8.4.1Two-State, Fourth-Order RungeKutta


Algorithm and Code

8.4 RungeKutta 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


difference approximation of the first derivative of the state
variable and the time:

g
f

f 3 = a 11 x1 (n) + 2 + a 12 x2 (n) + 2 + b1u T

2
2

vn +1 = vn +

vn
t
t

(8.19)

Although the Euler method is simple, it is inaccurate and


can be numerically unstable, meaning, the accumulated
errors in the response will eventually lead to the result of
positive or negative infinity. The accuracy of the Euler
method can be improved, by reducing the time step to improve the approximation of the derivatives. This has the
unintended effect, however, of increasing the round-off
error. A reduction in the number of time steps increases the
number of computations.
A more accurate and stable finite difference method is
the RungeKutta method, which resembles a Taylor series
expansion. The RungeKutta method is also known as a
quadrature method, a name which refers to the use of
graph paper with axes at 90 or at quadrature. The method was developed for manual evaluation, when no other
means of integration was possible. We will use a fixedtime step, fourth-order RungeKutta scheme. The term
fourth-order refers to the successive approximations
of each step that reduce the error to the fourth order. The
RungeKutta algorithm improves the accuracy of the forward step, by estimating half a step first, then using those
results to estimate a full step. The method uses the current
value of the state vector, x(n), as the operating point. An
intermediate value, x(n + T /2), is then calculated to project
forward by half a time step, T/2. This value is then used to
create the next intermediate value in a recursive formula.
The process is repeated. The intermediate values are then
weighted and summed to calculate a value for the state
vector, x(n +1).

The fourth-order RungeKutta algorithm is presented below


for a system of two-state equations, first in conventional
mathematical notation, then as MATLAB code. The Runge
Kutta algorithm replaces the Euler method in the for loop of
the preceding MATLAB examples:
f1 =
g1 =

(a
(a

)
x ( n) + b u ) T

x (n) + a 12 x2 (n) + b1u T

11 1

x (n) + a 22

21 1

f
g

f 2 = a 11 x1 (n) + 1 + a 12 x2 (n) + 1 + b1u T

2
2

f
g

g3 = a 21 x1 (n) + 2 + a 22 x2 (n) + 2 + b 2 u T

2
2
f4 =
g4 =

( a ( x ( n) + f ) + a ( x ( n) + g ) + b u ) T
( a ( x ( n) + f ) + a ( x ( n) + g ) + b u ) T
11

21

x1 (n + 1) = x1 (n) +

12

22

( f1 + 2 f 2 + 2 f3 + f 4 )
6

( g + 2 g 2 + 2 g3 + g 4 )
x (n + 1) = x (n) + 1
2

The corresponding MATLAB code follows.


%
% Runge-Kutta Difference Equation loop
% Two State Equations
%
for n = 1:N;
%
f1 = (a11*x1(n) + a12*x2(n) + b1*u)*T;
g1 = (a21*x1(n) + a22*x2(n) + b2*u)*T;
%
f2 = (a11*(x1(n)+f1/2) + a12*(x2(n)+g1/2) + b1*u)*T;
g2 = (a21*(x1(n)+f1/2) + a22*(x2(n)+g1/2) + b2*u)*T;
%
f3 = (a11*(x1(n)+f2/2) + a12*(x2(n)+g2/2) + b1*u)*T;
g3 = (a21*(x1(n)+f2/2) + a22*(x2(n)+g2/2) + b2*u)*T;
%
f4 = (a11*(x1(n)+f3) + a12*(x2(n)+g3) + b1*u)*T;
g4 = (a21*(x1(n)+f3) + a22*(x2(n)+g3) + b2*u)*T;
%
x1(n+1)= x1(n)+(f1+2*f2+2*f3+f4)/6;
x2(n+1)= x2(n)+(g1+2*g2+2*g3+g4)/6;
%
t(n+1) = n*T;
end;

8 Finite Difference Methods and MATLAB

480

f1 =
g1 =
h1 =

(a
(a
(a

x (n) + a 12 x2 (n) + a 13 x3 (n) + b1u T

11 1

)
x ( n) + b u ) T

x (n) + a 22 x2 (n) + a 23 x3 (n) + b 2 u T

21 1

x (n) + a 32 x2 (n) + a 33

31 1

h1

f
g

f 2 = a 11 x1 (n) + 1 + a 12 x2 (n) + 1 + a 13 x3 (n) + + b1u T

2
2
2

h1

f
g

g 2 = a 21 x1 (n) + 1 + a 22 x2 (n) + 1 + a 23 x3 (n) + + b 2 u T

2
2
2

h1

f
g

h 2 = a 31 x1 (n) + 1 + a 32 x2 (n) + 1 + a 33 x3 (n) + + b 3u T

2
2
2

h2

f
g

f 3 = a 11 x1 (n) + 2 + a 12 x2 (n) + 2 + a 13 x3 (n) + + b1u T

2
2
2

h2

f
g

g3 = a 21 x1 (n) + 2 + a 22 x2 (n) + 2 + a 23 x3 (n) + + b 2 u T

2
2
2

h2

f
g

h 3 = a 31 x1 (n) + 2 + a 32 x2 (n) + 2 + a 33 x3 (n) + + b 3u T

2
2
2

f4 =
g4 =
h4 =

(a
(a
(a

(
) )
( x ( n) + h ) + b u ) T
( x ( n) + h ) + b u ) T

11

( x1 (n) + f 3 ) + a 12 ( x2 (n) + g3 ) + a 13 x3 (n) + h 3 + b1u T

21

( x1 (n) + f 3 ) + a 22 ( x2 (n) + g3 ) + a 23

31

( x1 (n) + f 3 ) + a 32 ( x2 (n) + g3 ) + a 33

x1 (n + 1) = x1 (n) +
x2 (n + 1) = x2 (n) +
x3 (n + 1) = x3 (n) +

(f

+ 2 f 2 + 2 f3 + f 4 )

6
( g1 + 2 g 2 + 2 g3 + g 4 )

(h

6
1

+ 2h 2 + 2h 3 + h 4
6

8.4.2Three-State, Fourth-Order RungeKutta


Algorithm and Code
The RungeKutta 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 conventional mathematical notation for a set of three-state equations and the corresponding MATLAB code: Note the use of
ellipses (the three dots or periods) to continue a MATLAB
statement to the next line.
%
% Runge-Kutta Difference Equation loop
% Three State Equations
%
for n = 1:N;

)
%
f1 = (a11*x1(n) + a12*x2(n) + a13*x3(n) + b1*u)*T;
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;
%
f2 = (a11*(x1(n)+f1/2) + a12*(x2(n)+g1/2)
+ a13*(x3(n)+h1/2) + b1*u)*T;
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
%
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;

8.5 Programming Non-Linearities and Input Functions

20

Euler

The system equation relating the input force to the velocity of the mass is

Runge-Kutta

15

v1g

481

10
5

1 d F d 2 v b dv K
= 2 +
+ v
M dt
M dt M
dt

FK

Euler

Runge-Kutta

and the system equation relating the input force to the force
carried by the spring is

0
-5

N=2000

20

40

60

80

t, sec

d 2 FK
K
b d FK K
F=
+
+
FK
2
M
M dt
M
dt

100

The results produced by the RungeKutta and the Euler


methods, and the analytical solutions for the responses of the
state variables of a spring-mass-damper system are plotted in
Fig.8.9. The same time step was used for both RungeKutta
and Euler methods. The RungeKutta results and the analytical solutions are plotted on top of one another. The Euler
methods error increases with the number of steps.

Fig. 8.9 Comparison of the Euler, RungeKutta methods and analytical calculations. The RungeKutta results were indistinguishable from
the analytical results whereas the error between the Euler and analytical
results increased as the calculation progressed

h3 = (a31*(x1(n)+f2/2) + a32*(x2(n)+g2/2)
+ a33*(x3(n)+h2/2) + b3*u)*T;
%
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;

8.5Programming Non-Linearities and Input


Functions
The RungeKutta functions in Mathcad and MATLAB can
solve non-linear state equations if the non-linearities can be
expressed in a single, conventional mathematical expression.
The term conventional in this context means an expression
without a conditional or logical function, such as an if()
statement. For example, the power-law damping force of
Eq.8.16 and Fig. 8.10a is a non-linearity which can be described without a conditional statement. However, the approximation of the function created from three straight lines,
Fig.8.10b, requires conditional statements, Eq.8.20:

%
x1(n+1)= x1(n)+(f1+2*f2+2*f3+f4)/6;
x2(n+1)= x2(n)+(g1+2*g2+2*g3+g4)/6;
x3(n+1)= x3(n)+(h1+2*h2+2*h3+h4)/6;
%
t(n+1) = n*T;
end;

8 N sec
v1g if 6.4 v1g 6.4

6.4 m

27 8 N sec
(8.20)
Fb = 8 N +
v1g if 6.4 < v1g

10
6.4
m

27 8 N sec
v1g if v1g < 6.4
8 N

10 6.4 m

Compare the results of the Euler and RungeKutta finite


difference method solutions with the analytical result for
the state-space representation of a spring-mass-damper
system, where K = 1, M = 3, and b = 0.2. The input is
F (t ) = 10us (t ).

Fig. 8.10a Power law


damping, Eq.8.16.,
3
3

Fb = A0 /h 2 v1g v1g 2.

b Piece-wise continuous
approximation of the power
law damping superposed on the
trace of a with inflection points
at v1g = 6.4 m/sec, Fb = 8 N
and v1g = 6.4 m/sec, Fb = 8 N

b 27
20

20

Fb
0
N

Fb
0
N

-8

-20
-10

-5

m
v1g, ___
sec

10

-20
-27
-10

-6.4

-5

m
v1g, ___
sec

6.4

10

8 Finite Difference Methods and MATLAB

482
Fig. 8.11 a Saturation limits the
damper force, Eq.8.21. b Clipping truncates voltage measurement, Eq.8.21

b
20

Fb
0
N

v(t) 0
VDC

-20

-4

-2

-10

Fig. 8.12 a Bingham fluid with


a threshold pressure. b Deadzone of insufficient angular
velocity to transmit torque

-5

m
v1g, ___
sec

-6

10

t,sec

a
20

20

p12
0
MPa

T 0
N m

-20

-20

-0.01

-0.005

8.5.1 Common Non-Linearities


A non-linearity due to inflections points may be deliberately
created to meet design requirements. For example, when a
force exceeds a certain level, the resulting deflection may
bring the component into contact with a second helper
spring.
A non-linearity due to inflections points can also be the
result of exceeding the intended operating limits of a device.
For example, rather than a deflecting component contacting
another spring, it contacts the machine frame, changing its
bending mechanism.
A second type or class of common non-linearities is saturation or clipping. As the term clipping connotes, this
type of non-linearity is a result of a limit affecting the amplitude of a function, Fig.8.11a. An energetic device is saturated when additional power input does not yield additional
power output, or additional power output is negligible. For
example, an electronic amplifier is saturated when increasing the input signal does not increase the output signal. A ferromagnetic material is saturated when additional magnetomotive force, i.e., additional current through a winding, does
not increase the torque of the motor, or, more precisely, does
not increase the magnetic flux producing the torque above
that, which would exist without the ferromagnetic material.
Clipping has the same appearance as saturation but is due
to a limit of the acquisition or processing of data rather than a
power limitation in the physical phenomenon. For example,

m3
Q, ___
sec

0.005

0.01

-2

-1

rad
1g , ___
sec

a sinusoidal signal with an amplitude of 5.25 volts acquired


on an oscilloscope set to acquire up to five volts is clipped
at five volts, Fig.8.11b.


N sec
m

3.6 m v1g if v1g 7 sec

Fb =
v
m
25 N 1g if v1g > 7
sec
v1g

5.25 VDC sin(2t ) if v(t ) 5 VDC

v(t ) =
v(t )
5.0 VDC v(t ) if v(t ) > 5 VDC

(8.21)

(8.22)

Notice that the clipping described by Eq.8.22 is a function


of the output magnitude. The second condition in Eq.8.22
is sensible for a computer program but not for conventional
mathematics. It is an instruction to overwrite a variables
value based on the original value.
Backlash, due to clearance between the teeth of meshed
gears, slack, due to clearance between components in a force
load path or the lack of tension in belts, chains and cables,
and slip, due to insufficient shear force within a mechanism,
result in temporary lack of power transmission on startup or
loss of power transmission when a system reverses direction. Backlash and slack are manifested as dead-zones of
zero force or torque on either side of zero displacement when

8.5 Programming Non-Linearities and Input Functions


Fig. 8.13 a Check valve with a
threshold pressure below which
it does not open. b Self-locking
mechanism, which will only
rotate in the positive direction

483

a
20

20

p12
0
MPa

T 0
N m

-20

-20

-0.01

-0.005

the velocity has been zero. This requires integration of the


mechanism velocity to yield displacement as part of the finite difference code.
Thresholds are non-linearities, which manifest the opposite behavior of a dead-zone, Fig.8.12a. A threshold is a
level below which a finite force produces zero velocity. The
force must exceed the threshold level for the motion to occur.
Thresholds are found in all types of physical systems, either
by design or coincidence. Many electronic devices have voltage thresholds below which they will not conduct. A Bingham fluid is a fluid with a threshold due to shear strength.
It will not flow below its threshold pressure. E.C. Bingham
(18791945) was a professor of chemistry at Lafayette College who pioneered the study of rheology.
Directional dependency is often designed into components for functional requirements of the system, e.g., check
valves to prevent backflow and diodes limit current flow
to one direction, Fig.8.13. Diodes and many check valves
also have a threshold voltage or pressure below which there
is no flow. Mechanical systems can be locking or selflocking. In some cases such as ratchets, the locking mechanism permits motion only in one direction until the ratchet
is released. In other cases, such as a worm and worm gear,
the system can be driven in either direction by an input
torque on the worm but cannot be driven by output torque
on the gear. In addition to fluid non-linearities, the crosssectional area of a hydraulic piston/cylinders operating
rod reduces the piston area on that side. Consequently, the
maximum tensile force a double acting (push-pull) piston/
cylinder can produce is less than the maximum compressive force.
Directional dependency can also occur in translational
systems when either the operating range or the output exceeds the expected limits and components designed to be in
tension are placed in compression or components intended
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.

m3
Q, ___
sec

0.005

0.01

-2

-1

rad
1g , ___
sec

8.5.2Input Functions, Non-Linearities and the


RungeKutta Algorithm
The MATLAB code for non-linearities and input functions
must be inside the for-end loop which contains the Runge
Kutta algorithm, Fig. 8.14. Input functions are straightforward. Programming pulses into the RungeKutta algorithm
is easier than constructing a response function using superposition. The RungeKutta algorithm only needs the n1
values of the state variables, i.e., x1(n1), x2(n1), the values immediately preceding the step, and the nth value input.
The values of the state variables are saved as a matter of
course since the vectors of the values of the state variables is
the solution of the algorithm. All that must be programmed
is the value of the input.
Pulses are created by conditional statements. A single
pulse is created by an if statement before the RungeKutta
algorithm. For example, the following MATLAB code creates a pulse of amplitude 10 and a duration of 0.5sec.
%
for n = 1:N;
%
% Use if an if() to calculate the input u. Set u = 0 then
% overwrite with 10 if t < 0.5
u = 0;
if(n*T<=0.5);
u = 10;
end
% Beginning of Runge-Kutta Algorithm
f1 = (a11*x1(n) + a12*x2(n) + a13*x3(n) + b1*u)*T;
g1 = (a21*x1(n) + a22*x2(n) + a23*x3(n) + b2*u)*T;
A more complicated pulse requires more conditional statements and the form of an if-elseif statement.
%
for n = 1:N;
%
% Use if an if-elseif to calculate the input u. Set u = 0

8 Finite Difference Methods and MATLAB

484
Fig. 8.14 RungeKutta algorithm within a for-end loop. The
input value and state variable
dependent parameter values
are calculated at the top of the
loop, before the RungeKutta
algorithm. At the bottom of the
loop, after the state variables are
updated, velocities are integrated to displacements and state
variables are overwritten if they
exceed limits

Initialize variables
Calculate value of the input u.
Determine state variable dependent
parameter values.

for n=1:2000

f1=(a11*x1(n)+a12*x2(n)+a13*x3(n)+b1*u)*T;
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;
f2=(a11*(x1(n)+f1/2)+a12*(x2(n)+g1/2)+a13*(x3(n)+h1/2)+b1*u)*T;
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;

Runge
Kutta
Algorithm

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;
h3=(a31*(x1(n)+f2/2)+a32*(x2(n)+g2/2)+a33*(x3(n)+h2/2)+b3*u)*T;
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;

end;

Integrate velocities to displacements.


Test state variables with conditional
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
u = 0;
if(n*T <= 0.5);
u = 10;
elseif(n*T <= 1.0);
u = -10
end
% Beginning of Runge-Kutta Algorithm
f1 = (a11*x1(n) + a12*x2(n) + a13*x3(n) + b1*u)*T;
g1 = (a21*x1(n) + a22*x2(n) + a23*x3(n) + b2*u)*T;
Non-linear parameter functions that use the current value of
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
then the coefficients containing that parameter are calculated. In a linear RungeKutta calculation with constant parameters, the coefficients are calculated before the loop, since
they need only be calculated once.

8.5.3 Trapezoidal Integration


Integrating velocity to displacement is needed to impose
non-linearities created by geometric constraints. Trapezoidal
integration is the simplest, accurate method of numerical integration. As the name describes, the element added to the
integral is trapezoidal, Fig.8.15.

v1g(n+1) + v1g(n)
x1g = _____________ T
2

v1g (n)
m
___
sec
n-2

T n-1 T

T n+1

Step Number
Fig. 8.15 Trapezoidal integration of velocity v1g to displacement x1g.
The n+1 value of v1g was just calculated by the RungeKutta algorithm
and will be the n value when the for-end loop is iterated

Trapezoidal integration of state variables, which are velocities, occurs at the bottom of the for-end loop, after the
RungeKutta algorithm and the update of the state variables
from their x(n) values to their x(n+1) values.

Summary
Finite difference methods are necessary to solve non-linear
system equations. State equations are solved using finite
difference methods in all cases. The state-space representation is particularly convenient for non-linear dynamic
systems. The Euler method was the first method of finite
differences and remains the simplest. The RungeKutta

Problems

485

Flywheel
Rotational Inertia J
Compliant Shaft
Torsional Spring K
Fluid Coupling
Damping b
Angular Velocity
Input (t)
Fig. P8.1 A rotational mechanical system

method is the most widely used because of its accuracy and


numerical stability.

Problems
Problem 8.1 The rotational system shown schematically in
Fig.P8.1 comprises an angular velocity input driving a fluid
coupling attached to a compliant shaft modeled as a spring.
The compliant shaft drives a flywheel with mass moment
of inertia J. Two sets of parameters, Case I and Case II, are
tabulated.
Table P8.1 Parameter values
J

Case I

Case II

Case I

Case II

8.2.a Derive the state equations for this system.


8.2.b Derive the output equations for:
i The velocity of the mass M.
ii The force acting through spring K.
iii The angular velocity difference across spring K.
8.2.c Express the state and output equations in matrix form.

x,v

F(t)

Problem 8.2 A mass M supported on frictionless roller is


shown in Fig.P8.2a. A dashpot with damping b and a spring
with spring constant K are connected between the mass and
ground. The mass is acted upon by an applied force F(t).
The system has a linear and a non-linear configuration.
The linear configuration has linear dampers, which are described by the equation Fb=bv1g. The non-linear configuration has non-linear dampers designed so that they exert a
damping force when the damper is in compression but not
when the damper is in extension, as shown in Fig.P8.2b.
Table P8.2 Parameter values

8.1.a Derive the state equations for this system.


8.1.b Derive the output equations for:
i The angular velocity of the inertia J.
ii The torque acting through the compliant shaft
spring K.
iii The angular velocity difference across the compliant shaft spring K.

Fig. P8.2a Second-order


translational mechanical system.
b Non-linear damper, which
exerts a damping force only in
compression

iv The angular velocity difference across the fluid


coupling b.
8.1.c Express the state and output equations in matrix form.
8.1.d For Cases I and II, calculate the systems eigenvalues.
If the system is underdamped, determine the ideal,
undamped natural frequency n, the damped natural
frequency d, and the damping ratio .
8.1.e The system is de-energized when it is acted on by a
step input of 10rad/sec. Write a MATLAB program
for the Euler method to solve the state equations, and,
using the solution of the state equations, solve the output equations for Cases I and II. Plot the responses of
the output variables.
8.1.f Repeat part e using the RungeKutta algorithm.
8.1.g Plot the Euler method and RungeKutta algorithm
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?

20

Fb
N 0

M
K
Frictionless rollers

-20
-10

-5

m
v1g, ___
sec

10

8 Finite Difference Methods and MATLAB

486
R1 is a check valve.
Permits flow only in
this direction.

P(t)

R1

L
D

Pump

vent to
atmosphere

R1

R2

v(t)

Fluid
Reservoir

Fig. P8.3a Fluid circuit schematic

Fig. P8.4a Schematic of a RLC circuit with a diode D

2,000

p(t)
psi

1,000

10

R2

t, sec

20

30

Fig. P8.3b Input pressure pulse

8.2.d For the linear Cases I and II, calculate the systems
eigenvalues. If the system is underdamped, determine
the ideal, undamped natural frequency n, the damped
natural frequency d, and the damping ratio .
8.2.e Write a MATLAB program to solve the linear state
equations using the RungeKutta algorithm for a step
input of 10N. Assume that the system was de-energized at time t=0. Include the code to solve the output equations and plot their responses. Run your code
using the parameter values for Cases I and II.
8.2.f Save your MATLAB script to a different file name
and edit the script to include the non-linear behavior
of the dashpot. Run your program for the same input
and parameter values as part e.
Problem 8.3 The fluid system shown schematically in
Fig.8.3a consists of a fluid power unit, modeled as a pressure source, two fluid resistances R1=100 MPasec/m3
and R2=200 MPasec/m3, a long run of piping, modeled
as a fluid inertance, I=40107 kg/m4 and a fluid accumulator (capacitance), C=6108m3/Pa. The system will be
modeled twice, a linear and a non-linear model. In the linear
model, the fluid resistance R1 has a constant resistance and
allows flow in both directions. In the non-linear model, the
fluid resistance R1 is a check valve which only allows flow
in the direction shown. The fluid resistance in the forward
direction is the same as the linear case.
8.3.a Derive the state equations for this system.

8.3.b Derive the output equations for:


i The volume flow rate from the pressure source.
ii The pressure in the fluid accumulator.
iii The volume flow rate through the fluid inertance.
iv The pressure difference across the fluid inertance.
and check their units.
8.3.c Express the state and output equations in matrix form.
8.3.d Calculate the systems eigenvalues configured with
the linear fluid resistance R1. If the system is underdamped, determine the ideal, undamped natural frequency n, the damped natural frequency d, and the
damping ratio .
The system was de-energized before the pressure pulse plotted in Fig.P8.3b was applied to the system.
8.3.e Write a MATLAB program to solve the linear state
equations using the RungeKutta algorithm for the
pressure input shown in Fig.P8.3b. Assume that the
system was de-energized at time t=0. Include the code
to solve the output equations and plot their responses.
8.3.f Save your MATLAB script of part e under a different
file name and edit it to solve the non-linear state equations using the RungeKutta algorithm for the pressure input shown in Fig.P8.3b.
8.3.g Compare the pressure across the fluid inertance in the
linear and non-linear cases and calculate the maximum change in pressure due to the check valve.
Problem 8.4 The electric circuit shown in the schematic,
Fig. 8.4a, consists of a voltage source, two resistors,
R1 = 0.1W and R2 = 1W, a capacitor, C=20106F, an inductor, L=1 106H, and a diode D.
Diodes conduct current only in the direction that the
triangle of their symbol points Fig.8.4b. This is a diodes
forward direction. Diodes have a threshold voltage, which
varies with the material and design of the diode. A typical
silicon diode has threshold voltage of 0.63 VDC before it
conducts in the forward direction, Fig.8.4c. An actual diodes
voltagecurrent plot is similar to Fig.8.4c. The small increase in resistance to the forward conduction with increased

Problems

487

Fig. P8.4b Diode schematic


symbol showing forward direction (direction of current flow).
cDiode voltagecurrent plot

Anode = Positive
The band on physical
diode and the bar on
the schematic symbol
indicate the cathode.

+
iD
-

Cathode = Negative

10
5

v(t)
0
VDC
-5

0.05

0.1

0.15

t, msec

-10

Fig. P8.4d Input voltage

current 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 linear configuration without the diode, and a non-linear configuration with the diode.
8.4.a Derive the state equations for the linear configuration
and check their units.
8.4.b Derive output equations for the following variables of
the linear system.
i The voltage across capacitor C.
ii The current flowing through inductor L.
iii The voltage across inductor L.
iv The current through resistor R1.
v The current through resistor R2.
8.4.c Write the state equations and output equations in vector-matrix form.
8.4.d Calculate the systems eigenvalues If the system is
underdamped, determine the ideal, undamped natural
frequency n, the damped natural frequency d, and
the damping ratio .
The system is initially de-energized before the voltage pulse
shown in Fig.P8.4d was applied.
8.4.e Write a MATLAB program to solve the linear state
equations using the RungeKutta algorithm for the
voltage pulse input shown in Fig.P8.4d. Assume that
the system was de-energized at time t=0. Include
the code to solve the output equations and plot their
responses.
8.4.f Save your MATLAB script of part e under a different
file name and edit the script to solve the state equa-

30
20

iD
A 10
-

0
-0.5

0.5

vD , VDC

1.0

1.5

tions for the non-linear configuration, which includes


the diode, for the voltage input shown in Fig.P8.4d.
8.4.g C
 ompare the voltage across the inductance in the linear and non-linear systems and calculate the maximum change in voltage due to the diode.
Problem 8.5 Two rotational mechanical systems are shown
schematically in Figs.P8.5a and P8.5b. Both systems contain
a torque source, a gear with rotational inertia, J1, supported
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
gear, driven by a pinion with negligible inertia. In System II,
the gear is a worm gear, driven by a worm with negligible
inertia. The gear ratios of Systems I and II are the same: 20
rotations of the pinion, or worm, to one rotation of the gear.
Table P8.5 Parameter values for systems I and II
K
J1
J2
b1
0.5kg m 2

To be
To be
calculated calculated

10

N m sec
rad

b2
20

N m sec
rad

There is one significant difference between the two systems. The spur gears of System I can be driven by either
the torque source or the output shaft, K. However, the worm
gears of System II can only be driven by the torque source. A
worm drive is self-locking. It cannot be driven by its output shaft on the worm gear, only the input shaft on the worm.
The angular velocity of the pinion in System I need not
have the same sign as the torque applied by the torque source.
However, the angular velocity cannot have the opposite sign
of the torque from the torque source. The angular velocity of
the worm in System II can be of the same sign as the torque
applied by the torque source or zero.
The spur gears are counter-rotating. Consequently, the angular velocities of the two shafts have opposite signs. We are
free to define the signs for the rotation of the input and output
shafts of the worm drive. For convenience, we will define the
sense of rotation of the input and output shafts to have opposite signs to correspond to therotation of the spur gear system.
8.5.a Derive the state equations for Systems I and II.

8 Finite Difference Methods and MATLAB

488
Fig. P8.5a System I spur gear
drive

Bearings,Negligible
damping

Pinion, Negligible
Inertia

T(t)
Torque
Source

Load
Inertia J 2
Torsional Shaft
Spring Constant, K

Bearings
Damping b1

Fig. P8.5b System II worm


drive

Bearings
Negligible
damping

Worm Gear
Inertia J 1

Worm
Negligible
Inertia

Load
Inertia J2

Torsional Shaft
Spring Constant, K

Bearings
Damping b1

Bearings
Damping b 2

T(t)

Bearings
Damping b2

Gear
Inertia J1

Torque
Source

1 in

T(t)
N m

10

4 in 2 in
32 in 36 in

t, sec

Fig. P8.5d Input torque pulse

Section A-A

Fig. P8.5c Flywheel J2

8.5.b The shaft, K, is carbon steel, solid with a circular


cross-section, 2 in. in diameter and 12 ft long. Calculate its torsional spring constant, K, as described in
Chap.4.

8.5.c The flywheel, J2, shown in the plan and cross-section


in Fig. P8.5.c, is carbon steel. Calculate the mass
moment of inertia (rotational inertia), J2, using superposition, as described in Chap.4. The spokes are 1 in.
wide.
8.5.d Draw a flowchart of the logic needed in the state equation solver loop, to prevent backwards (positive)
rotation of the worm gear, and to lock the worm gear,
when the input torque is turned off.

Problems

489
x,v

K
F(t)

b2

b1

M1

F(t)
N

M2

10

20

t, sec

40

Fig. P8.6c Input force pulse


Fig. P8.6a Third-order mechanical system
Table P8.6 Parameter values
M1
M2

20

Fb2
N

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

-20

-10

-5

m
v2g, ___
sec

10

Fig. P8.6b Non-linear damper force, damper b2 Case I. The damping


coefficient in compression is that in TableP8.6. The damping coefficient in extension is one half of the tabulated value

The system is at rest and relaxed for t < 0, before an input


torque pulse with magnitude of 10 N m and a duration of
one second is applied at time, t=0, as shown in Fig.P8.5d.
8.5.e Write a MATLAB code to solve the state equations
of System I with the RungeKutta algorithm for the
pulse input shown in Fig.P8.3d. The duration of the
simulation should be 2sec. Plot the state variables.
8.5.f Modify the code to solve the state equations of System
II, again plotting the state variables.
8.5.g Calculate the maximum stress in the shafts of System
I and System II. If the yield strength of the steel is
60ksi, did either shaft fail?
Problem 8.6 A translational mechanical system contains
a spring, two masses, and two dampers acted upon by an
applied force is shown in Fig.P8.6a. The system has linear and non-linear configurations. There are two cases for
both configurations. The two cases for the linear configuration of the system use the parameters in TableP8.6. The
two cases of the non-linear configuration use the mass
and spring constant values from TableP8.6, but have nonlinear dampers which have half the damping coefficient
b in extension as they do in compression, as shown in
Fig. P8.6b. The values of damping coefficient b1 and b2
in TableP8.6 are the higher compression values for the
non-linear dampers.

8.6.a Derive the state equations for this system.


8.6.b Derive the output equations
i The velocity of mass M1.
ii The force acting through the spring K.
iii The velocity of mass M2.
iv The force acting through damper b1.
v The force acting through damper b2.
8.6.c Express the state and output equations in matrix form.
8.6.d For the linear configuration, calculate the systems
eigenvalues for Cases I and II. If the system is underdamped, determine the ideal, undamped natural frequency n, the damped natural frequency d, and the
damping ratio .
8.6.e For Cases I and II of the linear configuration, write a
MATLAB program to solve the state equations and
the output equations for the force pulse plotted in
Fig.P8.6c. Plot the responses of the output variables.
8.6.f Edit a copy of the MATLAB program of part e to solve
state and output equations of the non-linear configuration for Cases I and II. Plot the responses of the output
variable to the force pulse of Fig.P8.6c.
Problem 8.7 A translational mechanical system contains a
mass M sliding on a lubricating fluid film with damping b.
The mass is acted upon by an applied force F(t). Spring K1 is
attached between the mass and ground. Spring K2 is attached
to ground but is not attached to the mass. When the system is
at rest and relaxed, there is a gap of width d between Spring
K2 and the mass M. Spring K2 and is only in contact with the
mass when the position of the mass compresses spring K2.
Two linear and one non-linear configurations of the system will be investigated. The two linear configurations are
(1) Spring K2 is absent from the system, and (2) Spring K2
is attached to mass M. The non-linear configuration is that
shown in Fig.P8.7 with gap d between the mass and spring
K2 when the system is at rest and relaxed. We will investigate
two cases. (1) The gap d of the de-energized system equals

8 Finite Difference Methods and MATLAB

490

x,v

8.7.g Save your MATLAB script of part f under a different


name and edit it to solve the non-linear state equations
which include a gap between the mass and spring K2.
Solve the state and output equations for the input and
conditions of part f and the two different gap widths.

K2
F(t)

K1

Problem 8.8 The schematic of an electric circuit with a voltage source, two inductors, two capacitors, a resistor, and a
diode is shown in Fig.P8.8a.

Lubricating fluid
Damping b

Table 8.8a Parameter values

Fig. P8.7 Second-order translational mechanical system

L2

L1
D
C1

v(t)

C2

Fig. P8.8a A non-linear RLC circuit

zero, i.e. the spring is in contact with but not attached to the
mass, and (2) the gap equals half the steady-state displacement of the mass under the load of 1,000 N. There are two
sets of parameters, Case I and Case II.
Table P8.7 Parameter values
M

K1

K2

Case I

50

10,000

20,000

60

Case II

50

20,000

10,000

60

8.7.a Derive the state equations for this system using Kequiv
as the spring constant.
8.7.b Derive the output equations for:
i The velocity of the mass M.
ii The force acting through spring K.
8.7.c Express the state and output equations in matrix form.
8.7.d For the linear configuration, calculate the systems
eigenvalues for the parameter values of Cases I and
II. If the system is underdamped, determine the ideal,
undamped natural frequency n, the damped natural
frequency d, and the damping ratio .
8.7.f Write a MATLAB program to solve the linear state
equations using the RungeKutta algorithm for a step
input force F(t)=1,000 N us(t). Assume the system is
at rest and relaxed before the input acts on the system.
Solve the output equations for the parameter values of
Cases I and II and plot their responses.

C1

C2

L1

L2

1 F

2 F

3 H

4 H

2W

Diodes conduct current only in the direction that the triangle of their symbol points Fig.8.8b. This is a diodes forward direction. Diodes have a threshold voltage which varies with the material and design of the diode. A typical silicon
diode has threshold voltage of 0.63 VDC before it conducts
in the forward direction, Fig.8.8c. An actual diodes voltagecurrent plot is similar to Fig.8.8c. The small increase
in resistance to the forward conduction with increased current 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 linear configuration without the diode, and a non-linear configuration with the diode.
8.8.a Derive the state equations for the linear system and
check their units.
8.8.b Derive output equations for the following variables of
the linear system.
i The voltage across capacitor C1.
ii The voltage across capacitor C2.
iii The current through inductor L1.
iv The current through inductor L2.
v The voltage across inductor L1.
vi The voltage across inductor L2.
vii The current through resistor R.
8.8.c Write the state equations and output equations in vector-matrix form.
8.8.d Calculate the linear systems eigenvalues. If the system is underdamped, determine the ideal, undamped
natural frequency n, the damped natural frequency
d, and the damping ratio .
The system is initially de-energized before the voltage
pulse shown in Fig.P8.8d was applied.
8.8.e Write a MATLAB program to solve the linear state
equations using the RungeKutta algorithm for the
voltage pulse input shown in Fig.P8.8d. Include
the code to solve the output equations and plot their
responses.

Problems

491

Fig. P8.8b Diode schematic


symbol showing forward direction (direction of current flow).
cDiode voltagecurrent plot

Anode = Positive
The band on physical
diode and the bar on
the schematic symbol
indicate the cathode.

+
iD
-

Cathode = Negative

30
20

iD
A 10
0

-0.5

10

0.5

vD , VDC

-5

1.5

x
1

1.0

y,v

v(t)
0
VDC

t, millisec

-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

Youngs modulus E
Density
Width b

M
L

Fig. P8.9a Linear cantilevered beam with a cylindrical mass attached


at distance L from the support

8.8.f Save your MATLAB script of part e under a different


file name and edit it to solve the non-linear state equations, which include the diode in the circuit, for the
voltage input shown in Fig.P8.8d.
8.8.g Compare the voltage across the inductances in the
linear and non-linear systems and calculate the maximum change in voltage due to the diode.
Problem 8.9 A cantilevered beam with rectangular crosssection is shown in Fig.P8.9a. Attached to the beam is a
cylinder of mass M. The diameter of the cylinder equals the
width of the beam w. The height of the cylinder is twice its
diameter.
8.9.a Determine the undamped angular frequency of the
beam and mass system if they are carbon steel and
L=50 mm, w=5 mm, and h=1 mm.

8.9.b Determine the damping coefficient b needed for the


system to have a damping ratio of =0.04, using the
velocity the point on the beam at distance L from the
wall as the location of the velocity node 1. Determine
the eigenvalues of the system.
8.9.c The design is modified adding a support such that
there is a gap d=1mm below the bottom surface of
the beam, when the system is at rest and relaxed, at the
distance of x=30mm from the support, Fig.P8.9b.
When the beam is in contact with the support during
downward deflection, approximate it as a cantilevered
beam 20mm long. The spring constant of the beam
is greatly increased. Conversely, the effective mass of
the beam is reduced when beam is in contact with the
support.
The deflection equation for a point at distance x from the
support of a cantilevered beam of length L is

x =

Fx 2
(3L x )
6 EI

The area moment of inertia of a rectangular cross-section


about a horizontal line through its centroid is
I=

wh3
12

8 Finite Difference Methods and MATLAB

492
Fig. P8.10a Electromechanical
system with a compliant shaft
belt slip

Voltage Source v(t)

Electrical Resistance R

Compliant Shaft
Torsion Spring K1
Flywheel 1
Rotational Inertia J 1
Bearings Damping b1

DC Electric Motor
Torque constant K T
Flywheel 2
Rotational Inertia J 2
Bearings Damping b 2

6 in. Diameter Pulley


Belt

4 in. Diameter Pulley

Fig. P8.10b The compliant


belts taut and slack sides switch
positions during an oscillation.

vsurface

Slack
1g

K2

n+1

Taut
2g

Taut

Driving
Pulley

n-1

vsurface

Driven
Pulley

vsurface

g
Fig. P8.10c Linear graph representing a compliant belt as a translational spring between two rotational to translational transducer interfaces.

Determine the effective mass and stiffness of the beam with


the support as a function of the vertical displacement y of
node 1, at the distance L from the wall.
8.9.d Derive the state equations for this system.
8.9.e Derive the output equations for:
i The velocity of node 1.
ii The force acting through the equivalent spring K.
iii The angular velocity difference across spring K.
8.9.f Express the state and output equations in matrix form.
8.9.g Write a MATLAB program to solve the linear state
equations using the RungeKutta algorithm for a step
input of F0=10N. Assume that the system was deenergized at time t=0. Solve the output equations and
plot the responses of the output variables.
8.9.h Save your MATLAB script with a different file name
and edit the script to include the effect of the support

Slack

Driving
Pulley

vsurface

2g

Driven
Pulley

positioned under the beam. Run your program for the


same step input of F0=10N.

n+2

1g

Problem 8.10 An electromechanical system is shown in


Fig.P8.10a. A voltage source drives a DC motor, which has
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
inertia J1. Flywheel1 is supported on bearings with damping
b1. Flywheel2 has mass moment of inertia J2 bearings with
damping b2, and is belt driven. A belt runs between flywheels
one and two. The belt slips on the 4-in. diameter pulley when
the force carried by the belt equals 20 N, limiting the maximum force in the belt to 20 N.
Table P8.10 Parameter values
R

KT

J2

b2

5 kgm2

Nm
A
N m sec
rad

K1
800

J1
Nm
rad

b1

6 kgm2 1 N m sec
rad

K2
8,000

Nm
rad

The system model without belt slip is the linear model.


The system model with belt slip is the non-linear model.
8.10.a Derive the state equations of the linear model and
check their units.

Chapter 8 Appendix

8.10.b Derive output equations for the following variables:


i The current drawn from the voltage source.
ii The torque acting through the compliant shaft K1.
iii The angular velocity of flywheel J1.
iv The angular velocity flywheel J2.
8.10.c Write the state equations and output equations in
vector-matrix form.
8.10.d Calculate the systems eigenvalues for the linear
case with no belt slip. If the system is underdamped,
determine the ideal, undamped natural frequency n,
the damped natural frequency d, and the damping
ratio .
The system was de-energized before a step input of 48 VDC
was applied.
8.10.e Write MATLAB code to solve the state equations and
output equations of the linear model with the Runge
Kutta algorithm for the step input of 48 VDC. Plot
the responses of the output variables.
8.10.f Save the MATLAB script of part e under a different name and edit the code to solve the state equations and output equations of the non-linear model
for the step input of 48 VDC. Plot the responses of
the output.

Chapter 8 Appendix
Introduction to Programming and MATLAB
Why study computer programming, if engineering software
is available? Further, the sophisticated computer programs
used in engineering design and analysis are enormous. They
often required teams of programmers and years to develop.
What can an individual engineer and novice computer programmer accomplish of value? Fortunately, one does not
need to be an expert programmer to write useful computer
programs. A little knowledge of computer programming
goes a long way.
Competent engineers must be able to write simple computer programs for data acquisition and reduction, as well
as design computations. It is helpful to know some common
programming syntax and techniques, since they appear in
many user interfaces. An understanding of the fundamentals
of computer programming is important to maximize the utility of most engineering software. Most engineering software
packages allow the user to write either macros or scripts,
which are short programs, to automate repetitive processes
or iterative computations.
Computers are not only an important tool in the design
process. Microprocessors are now widely used as embedded controllers in machines. Microprocessors are microcomputers which contain the central processing unit (or
CPU which performs arithmetic and logical operations on
data), memory, and input and output devices on a single chip.

493

Now that microprocessors capable of machine control cost


less than one dollar, only the simplest machines are purely
mechanical. A machines functionality can be enhanced, and
its performance improved, by microprocessor control. A machine designer must understand the basics of how microprocessors function, to understand the machine they are part of.

A Brief History and Classification of Computer


Programming Languages
Computer programming languages have evolved to meet the
changing demands of users. The very first computers were
programmed in a machine-level language, which specified
how to implement each individual step required to manipulate the data. Machine-level code, or instructions, are at
the level of detail needed to describe a complicated handheld calculator computation, i.e., recall x, recall y, add x and
y, store the result in z, etc. The instructions, however, are in
binary code, since the computers digital circuitry functions
with 0s and 1s. The computer machine code equivalent to
a calculators recall is move data, which is the operation of reading data at a location in memory, and writing it
to a different location. Typical machine code for the operation, move data, is the four-digit binary code, 1011. Binary
code is difficult for humans to work with. The codes, 1101,
1011, and 1110, are different instructions. It was not long
before the second-generation computer languages, called
assembly languages, were created. Assembly languages
use an alphanumeric code in place of binary code, making
them much easier for humans to understand. The alphanumeric code incorporates three-letter abbreviations for the
machines operations. For example, MOV is the assembly
code for move data and MUL is the assembly code for
multiply. The most important memory locations, known as
registers, are given alphanumeric codes, such as AX and B1.
Octal (base8) or hexadecimal (base 16) can be used in place
of their 8, 16, or 32 binary-digit addresses.
Assembly language code is still commonly used to program the small microprocessors used for machine control. It
was superseded for general use in the early 1950s, however,
when FORTRAN (an acronym for FORmula TRANslation),
the first of the third-generation computer languages, was
introduced. FORTRAN was the dominant program for engineering computations through the mid-1970s. FORTRAN
is now known as a procedural language. The syntax and
structure of a procedural computer language are oriented
towards algorithm, i.e., the logic of the computation. Early
procedural languages include ALGOL (ALGOrithmic Language, mid-1950s), COBOL (COmmon Business Operations
Language, late 1950s), LISP (LISt Processing language,
early 1960s), and BASIC (Beginners All-purpose Symbolic
Instruction Code, mid-1960s). BASIC was implemented as
an interpreted, rather than a compiled, language. An

494

interpreted language is converted into machine code line


by line, as the program executes. Compiled languages are
converted into machine code, before the program executes.
Compiled programs run significantly faster than interpreted
programs. Procedural languages remain sufficient for programming engineering computations. FORTRAN was last
revised in 2008. LISP is used in Autocad. The current manifestation of BASIC is Microsofts Visual Basic 2013.
Two diametrically opposed trends impacted the development of computer languages in the late 1960s. Computer
programs increased in size, while computers themselves
decreased in size. Computer scientists found that some aspects of early procedural languages became unmanageable, as programs grew in size and were written by teams
of programmers. Pascal (mid-1970s) was the first of the new
structured programming languages. Structured programming languages modularized a large programming task into a
set of smaller tasks, and provided a structure for the modules
to interact. Modularizing the program code also allowed for
easier reuse of portions of the code in other programs. Simultaneously, minicomputers using newly developed integrated electronics provided low-cost competition to large and
expensive mainframes. These smaller computers prompted
a minimalist approach to developing applications to run on
them. The language C (early 1970s) is a procedural language
with an enriched set of operators. C also provides access to
bit-level operations, formerly only possible using assembly
language. C remains an important language today. Many microprocessors programmable in assembly language are also
programmable in C. Although the richness of C permits arcane programming practices, it is also possible to program in
C in a style close to FORTRAN, BASIC, or MATLAB. C is
an open source language, meaning it is available for free.
Current programming languages are object-oriented,
as opposed to the earlier procedural-oriented programming
languages. The intent of object-oriented programming is to
create modular and encapsulated code, where encapsulation means that the code can safely be used in a modern Windows-based program with millions of lines of code,
without causing conflict with other portions of the program.
In the abstract, the modern, enormous computer programs
can be written in the early procedural languages. In practice,
however, developing and debugging gigantic programs written in procedural languages would result in software overpriced and obsolete, by the time it went to market.
Object-oriented programming languages focus on the
data, not on algorithms. At first, this sounds impossible. Of
what use are data without algorithms to manipulate and interpret them? Object-oriented programming languages package the algorithms needed into a module called a class,
specific to each type of data, or object, in the program.
A familiar example of a data type is a complex variable.
Complex variables are manipulated differently from real

8 Finite Difference Methods and MATLAB

variables. The addition operator for real variables is part of


the programming language. The algorithms to create an addition operator for complex variables would be part of the
class, Complex. The class, Complex, would include other
properties, such as calculation of the magnitude, and operations, such as multiplication and division of complex numbers. The class, Complex, would be added to any program
which includes complex variables. Each complex variable
used in the program would be an object.
The first widely adopted object-oriented programming
language was C++ (early 1980s). C++ was developed from C.
C programs can run in most implementations of C++. Today,
C++ seems to be the dominant language used in academic
research. Microsoft gives away stripped-down versions of
C++, and its modernized version, called C#, with no multicore support or 64-bit math. C# has simpler syntax than C++.
Unfortunately, object-oriented languages are more difficult to
learn than procedural languages. There is twice as much to
understand. Object-oriented languages have all of the syntax
and operators for their procedural aspect, plus as much again
in syntax and operators for their object-oriented aspect.
We will work with the programming language, MATLAB. MATLAB is a high-level programming language
developed for engineering and scientific applications. High
level means that many of the tasks, which must be programmed in other languages, are performed automatically
in MATLAB. Specifically, all of the mathematical functions
you can imagine, along with an engineer-friendly plotting
function, are provided. MATLAB is an interpreted not compiled language. It runs slowly, compared to C or C++. Further, an academic license limits MATLAB in ways which
can affect the accuracy of computations. However, programs
will be fairly simple, as well as run quickly and accurately
enough for our purposes.

Fundamentals of Procedural Programming


One purpose in learning procedural programming is to automate numerical methods that were developed in the late
1800s and early 1900s, before the invention of the digital
computer. The algorithms were initially implemented using
desk calculators, which could add and multiply. The early
desk calculators were purely mechanical, crank-operated devices, which later evolved into electromechanical designs.
They were enormously complex and quite expensive. The
cost and complexity of these machines stimulated the design
of purely electronic calculators, which led to the invention of
the microprocessor.
Computer programs are instructions coded in the syntax
specific to a particular computer language. Dozens of computer languages exist. Different computer languages were
written to provide convenient methods of performing certain

Chapter 8 Appendix

495

tasks, such as creating attractive graphical user interfaces,


manipulating strings of characters, or performing numerical calculations. Although computer languages can be very
different, there are fundamental, logical, and mathematical
entities and operations which are common to all computer
languages.
Three fundamental functions must be performed by a
computer to execute an engineering algorithm.
1. Store and retrieve data.
2. Evaluate conditional statements comparing the values of
two variables.
3. Perform arithmetic operations.
We will focus on these fundamental aspects of programming. Our development will begin with a brief overview of
the physical and logical design of data storage. We will then
introduce flowcharts as a technique for graphical presentation of programming logic. Our first programming will use
the TI-89 calculator. This is historically appropriate, since
Texas Instrument introduced some important mid-scale integrated circuits in the late 1960s. We will then move to
MATLAB, a powerful high-level language. Finally, we will
compare MATLAB and the programming language, C.

A Brief History of Computer Memory


Data are stored in the computers memory, exploiting a physical phenomenon or device, which can exist in one of two
states (such as off or on; low or high; open or closed; deenergized or energized; etc.), associated with the binary digits, 0 and 1. The earliest digital computers built during World
War II used vacuum tubes, or electromechanical relays as
memory elements, where the state of the device represented
a binary digit, or bit. Vacuum tubes and electromechanical
relays were eventually replaced by reed relays, which are
still commercially available. Reed relays are pairs of very
thin steel cantilevers sealed in tiny glass vials. A coil around
the vial closes the contacts. Vacuum tubes consume a great
deal of power with lower reliability than reed relays, but they
transition, or change states, faster. Memory devices after
World War II included delay lines, devices which use the
propagation time of a wave through a media to store information, such as an acoustic wave through mercury; and drum
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.

Discrete (as opposed to integrated) transistors and magnetic core memory took the place of vacuum tubes in the
1950s. Core memory was the dominant form of computer
memory in the late 1950s through the early 1970s, and is
still used in a few military and aerospace applications for
its reliability in extreme conditions. Core memory is a rectangular grid of orthogonal wires with tiny rings, or cores,
of a ferromagnetic material at the intersections. Currents
through two wires magnetize the core at their intersection in either a clockwise or counter-clockwise direction,
depending on the direction of the currents. The direction
of magnetic flux in the core is defined as 0 or 1. Discrete
memory was replaced by integrated circuits, beginning in
the late 1960s. The density of integrated circuits, as measured by the number of transistors per unit area, has approximately doubled every two years.

Bits, Bytes, and Words


We now consider the basic structure of computer memory.
The smallest element of computer memory is a single BInary digiT, a bit. Groups of eight bits are called bytes. Bytes
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
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
word comprising two 8-bit bytes, which is the size of words
in many mid-level microprocessors used in machine control.
Further, we will only discuss microprocessors to focus on
the basics, and avoid the complexity of the clever addressing
schemes used by computers.
In order to store data in and retrieve data from a computers memory, there must be an address, a unique name, for
each word. The number of words, which can be addressed
using N binary bits, equals the maximum value of binary
number of N bitsin other words, a binary number of N bits,
each of which is 1
Number of Addresses =

n = N 1

Example: The number of words which can be addressed with


an 8-bit word is:

Number of Addresses = 111111112


=

n = 8 1

2n = 2 N 1

n=0

= 20 + 21 + 22 + 23 + 24 + 25 + 26 + 27

n=0

= 1 + 2 + 4 + 8 + 16 + 32 + 64 + 128 = 255 = 2 N 1

8 Finite Difference Methods and MATLAB

496

The number of words which can be addressed by a 16-bit binary number is 216 1 = 65, 536 1 = 65, 53510. Sixteen-digit
binary numbers are difficult to work with. The hexadecimal
equivalent is used instead.
11111111111111112 = FFFF16
A common notation for addressing a bit within a byte or
a word is to number the bits from the least significant bit
(LSB), to most significant bit (MSB). For example, the bit
addresses in a 16-bit word range from 0 for the LSB, to F for
the MSB expressed in hexadecimal.
The bit number is appended to the word address after a
dot. For example, the address of the MSB in the 16-bit word
with the address, 352A, is 352A.F. The bit address, 352A.F,
is not a binary fraction. It is an address code. The dot is not
a radix. It is just a dot.

Binary, Octal, and Hexadecimal Numbers


The base of a number system is the number exponentiated
to yield the place values of that system. Decimal is base 10.
Binary is base 2. We will also work with octal, which is base
8, and hexadecimal, which is base 16. We write numbers using
positional notation, which we commonly refer to as place
value. In a decimal number, the decimal point, or radix, marks
the division between place values with exponents less than
zero, and place values with exponents greater or equal to zero
16-Bit Word
F

is not an exact decimal fraction, since it cannot be expressed


as a finite sum of natural decimal fractions.
Positional notation is used within all number systems,
including binary. Although the term, binary point, is commonly used in the United States, the proper term is radix.
Integer binary place values are to the left of the radix. Natural binary fraction place values are to the right, just as with
decimal numbers. Natural binary fractions and their decimal
equivalents are familiar, because US customary units use
fractions which are powers of one half.
Table A8.2 Binary place values from 24 to 25
25
24
23
22
21 20 21
100000 10000 1000 100 10 1 0.1
32
16
8
4
2 1 1
2
32

16

22 23
24
0.01 0.001 0.0001
1
1
1
16
4
8

0.5 0.25 0.125 0.0625

The place values for the first 16 binary digits to the left of
the radix, expressed as decimal numbers, are
Table A8.3 Binary place values from 215 to 20
215
214
213
212
211
210
32,768 16,384 8,192 4,096 2,048 1,024
27
26
25
24
23
22
128
64
32
16
8
4

29
512
21
2

28
256
20
1

Octal (base 8) was used to address memory in early computers which had very little memory. Programmable logic
controllers (PLCs) still use octal (base 8) for addressing
memory, because they need relatively little.
Table A8.4 Octal place values from 87 to 80
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

64

Table A8.1 Decimal place values, including natural decimal fractions


105
104
103
102 101 100 101 102 103
104
1
1
1
1
100,000 10,000 1,000 100 10 1
10 100 1,000 10,000

Hexadecimal (base 16) is used in modern computers,


because it allows large numbers to be represented easily,
and converts easily and exactly to binary. The decimal
number system has 10 digit symbols, zero through nine,
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,
A through F.

8-Bit Byte

8-Bit Byte

100,000 10,000 1,000 100 10

0.1

0.01 0.001 0.0001

Natural decimal fractions are the place values to the


right of the radix. The only decimal fractions which are exact
are those which can be represented by sums of the natural
decimal fractions. For example,
1 5
1 1 1 1 1
=
= + + + + = 0.5
2 10 10 10 10 10 10

262,144

32,768

4,096

512

Table A8.5 Hexadecimal digits


Decimal

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

Hexadecimal 0 1 2 3 4 5 6 7 8 9 A

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

10

Chapter 8 Appendix

497

Base Conversion

Yielding the conversion from decimal to binary:

Base conversion allows one to represent a given number in


different number systems. It can always be accomplished by
performing successive division, starting with the largest divisible place value, and working down in place value. The
process is repeated on the remainder of the each division,
until the conversion is carried to the precision desired.
We will illustrate the technique by converting the decimal
number, 47.8, into binary number, using the binary place values from 25 to 24. The decimal number, 32 = 2 5 , is divisible,
leaving a remainder of 15.8. The next binary place value,
16 = 2 4 , is not divisible. The process is repeated.


 



 



 










Binary Place Values 25

24

23

22

21

Dividend
Divisor
Division possible?
Binary Digits
Remainder
Binary Place Values
Dividend
Divisor
Division possible?
Binary Digits
Remainder

15.8
16
No
0
15.8
21
0.8
0.5
Yes
1
0.3

15.8
8
Yes
1
7.8
22
0.3
0.25
Yes
1
0.05

7.8
4
Yes
1
3.8
23
0.05
0.125
No
0
0.05

3.8
2
Yes
1
1.8
24
0.05
0.0625
No
0
0.05



 









 













Perform successive division on the number to be converted to octal, 97810

Base 10 Divisors
Value of 97810 in Octal

84
4,096
0

83
512
1

82
64
7

81
8
2

80
1
2

The result is 97810=17228. The method of successive division always works, but is not necessary when one is converting between number systems that are powers of a common
base, such as binary (base 2=21), octal (base 8=23), or hexadecimal (base 16=24). There is a shortcut.
Numbers can be converted among binary, octal, and hexadecimal by exploiting the power relationship between the bases.
Compare the place values of binary, octal, and hexadecimal:

%LQDU\3ODFH9DOXHV









Note the error of 0.0510 between the binary and decimal


numbers, even though the base conversion was to four binary places to the right of the radix. It is always possible
to perform an exact conversion of integer values between
bases. However, it is often impossible to convert fraction
values between bases exactly.
A second example: Convert 978 (base 10) to octal (base 8).
First, determine the octal place values in base 10 to a sufficiently large value to perform the division.


 
 HWF

The results are


47.8
32
Yes
1
15.8
20
1.8
1
Yes
1
0.8

47.810 = 101111.11002

2FWDO3ODFH9DOXHV












+H[DGHFLPDO3ODFH9DOXHV






















8 Finite Difference Methods and MATLAB

498

The relationship between the nth octal place value and the
corresponding binary place value is 8n=(2n)3. Similarly,
the relationship between the nth hexadecimal place value,
and the binary place value is 16n=(2n)4. Consequently,
it requires three binary digits, or places, to represent an
octal number, and four binary digits to represent a hexadecimal number. The shortcut is to convert the number to
the common base, binary in this case, and then to either
octal or hexadecimal base, by separating the binary number into groups of three or four binary digits, respectively,
and converting the binary group into a single digit in the
higher base.
Example. Express the decimal number, 935, in binary,
octal, and hexadecimal. Base 10 does not share a common
base with base 2, base 8, or base 16. The conversion from
decimal to binary must be accomplished by successive division, as above. Summarizing the result, the binary value of
93510 is




































Summarize the decimal, binary, octal, and hexadecimal


values:
93510=11101001112=16478=3A716

Computational Error on Conversion from Natural


Decimal to Natural Binary Fractions
As we have seen, integer values can be converted between
different bases exactly. However, fractional values may be
impossible to express exactly within a base. If they can be
expressed exactly within one base, we may not be able to
express them in a second base. We touched on the computational error inherent in conversion between decimal and
natural binary fractions above. It is an important topic, which
deserves elaboration.





 

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


decimal fractions, if enough significant figures are used.
Example:

Clearly, simply summing the binary fraction approximation


of a decimal fraction is an unacceptable timing algorithm.
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 natural decimal fractions cannot be expressed exactly as binary
fractions! For example, we can approximate 1
10 10 as
1
1
1
1
1
1
16 + 32 < 10 < 16 + 32 + 64
10
10
10
10
10
10
0.0001102 < 0.110 < 0.0001112
0.0937510 < 0.110 < 0.10937510
Note additional binary places do not solve the problem. Also,
binary fractions may appear smaller than their corresponding
decimal fractions, but they are not.
Example. Suppose you have a timer, timing by one hundredth of a second by summing the equivalent binary fraction. How much error would you accumulate in an hour?
First, we have to choose our binary fraction approximation
of

1
10010

1
1
1
1
<
<
+
12810 10010 12810 25610
0.007812510 < 0.0110 < 0.01171875010
0.0000000102 < 0.0110 < 0.000000112
Both available approximations have substantial error:
0.007812510 0.0110 = 0.00218750010
0.01171875010 0.0110 = 0.00171875010

Using the approximation with less error per count,


0.0110 0.000000112 , in one hour, the accumulated error is

ignorance of the magnitude of the error, in converting between natural decimal fractions and natural binary fractions.
These errors could have been prevented by using natural binary fractions, which can be converted to decimal fractions
exactly.

Data Types
The binary numbers used for word addresses are positive binary integers. How does a computer store negative binary integers or real binary numbers, i.e., binary fractions? Also,
how are large numbers, those greater than 2N1, stored?
They are coded, and in order for a computer to manipulate
numerical data, it must know the coding, known as the data
type, associated with the data, in a specific word (or words)
in memory. Some languages will assume a default data type,
if the programmer does not declare a variable to be a certain data type. Other languages will not compile a program,
if all variables used are not declared and assigned initial values. The discussion that follows is general, since the specific
term used for a generic data type, and the number of bits
associated with it, varies with both computer and programming languages.

Integer and Signed Integer Variables


The simplest data type is a positive binary integer. No coding is needed, other than to associate place value with location within a word, since increasing numbers of bytes can be
used to store integers. Various adjectives and acronyms are
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.)
The signed integer data type requires coding to distinguish 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

8 Finite Difference Methods and MATLAB

500

(MSB) to 1, to indicate a negative number. Negative numbers are stored in twos complement form. Twos complement form is created in two steps. First, every bit in the number, except the sign bit, is inverted. If a bit is 0, then it is
changed to 1. If a bit is 1, then it is changed to 0. Next, the
number, 1, is added to the number.
Example. Create the twos complement of the binary
number, 1101011, by using a 16-bit word.
Sign bit (MSB) set to 1, followed by the binary number,
1101011:
1000000001101011
Bits 0 through E are inverted:
1111111110010100
One is added to create the twos complement:
1111111110010101
We will test the usefulness of twos complement by performing the addition, 1101011+1101011, which should yield
zero:
1111111110010101 Two s complement of 1101011
+ 0000000001101011
0000000000000000

Floating Point Variables


Floating point is the term used by computer scientists for
scientific notation, i.e., the product of a decimal fraction between 1 and 10, and a multiplier which is a power of 10.
Varying the exponent has the effect of moving the radix or
decimal point, hence the term floating point. The coding of
floating point numbers is more involved than for signed integers, since both the mantissa and the exponent are signed.
We will not delve into details, other than to note that more
bits are allotted to the mantissa than the exponent. For example, if the floating point data type was 4 bytes or 32 bits
long, the mantissa may be 24 bits and the exponent 8 bits.
Double is a common name for the data type with twice, or
double, the internal storage and, hence, numerical precision
of the floating point data type.
Computer code is restricted to the characters of a keyboard. Consequently, the product of the decimal fraction
multiplied by the power of 10 is expressed, omitting the base
of 10 and preceding the exponent of 10 with E (or e) for
Exponent, e.g.,
352.478 3.52478 102 3.52478E02

Boolean or Logical Variables


Boolean variables, also known as logical variables are a
late addition to programming languages. They are unnecessary, since one can use integer variables with values of either
0 or 1 in their place. They were added to programming languages to prevent errors that result from integer variables,
which were to be restricted to the two Boolean states, being
accidently assigned other values.

Procedural Logic and Flowcharts


The numerical methods used in engineering can be very
sophisticated. Some required decades to reach their current
form. Fortunately, writing a computer program to implement an existing numerical method is far less of a challenge.
Most of the effort is spent in developing an understanding of
the numerical method, before the program is written, then
searching for any coding errors afterwards.
Flowcharts are a graphical representation of the sequence,
in which the instructions of a program are to be executed.
Although not all types of programs can be depicted in a flow
chart, our programs can. A flowchart presents the logic of
a program, without the details of the specific programming
language syntax. Flowcharts serve the same function for
computer programs, as outlines serve for papers. Flowcharts
show the flow of logical control through an algorithm.
They consist of blocks, which contain instructions or operations, and arrows connecting the blocks to indicate the
flow of logical control.
Rectangular blocks are executable blocks, which represent actions or calculations. The action or calculation may be
a single, simple step, or it may be many complicated steps.
The action or calculation may be vaguely described in English or expressed explicitly in mathematics. There can only
be one logical path entering the block and only one logical
path exiting the block. The direction of the logical flow is
indicated by arrows.
Diamond-shaped blocks are decision blocks, per
Fig.A8.3. They contain conditional statements, i.e., greater
than, less than, equal to; or a Boolean or Logical variable;
or a result which must be either True or False. There can
only be onelogical path into the decision block. There are
two logical paths which exit. Each path must be identified by
the logical result or value it represents, i.e., Yes or No; True
or False; etc. The logical flow branches at decision blocks.
The exit from the decision block taken by the logical control
is governed by the result of evaluating the conditional statement, value of the Boolean variable, or logical result. A diamond shape is standard for a decision block, but that shape
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

x < 0.001 OR y < 0.2 ?

True

False
Fig. A8.3 The standard
diamond-shaped decision block

Fig. A8.4 An alternative, non-standard, decision block

No

n > 10 ?

Yes

Fig. A8.5 A logical


loop. Control exits
the loop, when the
condition expressed in
the diamond-shaped
decision block is
satisfied

Enter Loop

Instructions for
actions or computations

use, as needed, a non-standard, alternative style, decision


block, per Fig.A8.4.

Logic Loop
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
calculations. The instructions within the loop are executed
repeatedly, either for a fixed number of times, or until a logical condition is satisfied. If the programmer errs and neglects
to create an exit condition, or creates an exit condition that
is impossible to satisfy, then the logic is trapped in an endless loop.

Nested Loops
It is common to use nested loops with one loop inside another, Fig.A8.6. Each loop has a counter variable. The inner
loops counter variable must complete its index through its
range for the outer loops counter variable to index once
nested loop are useful for working with arrays and matrices.

Flowchart Rules and Guidelines


Although an international standard for flowcharts exists,
there is no reason to abide by it, when drawing flowcharts for

No

Done ?

Yes
Exit Loop

ones 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
decision (diamond) blocks which have Yes and No or
True and False paths.
2. Branching can only occur at a decision block.
The following are a few guidelines which will eliminate errors and improve the readability of flowcharts.
1. Every block in the diagram should be on at least one logical path from the Start block to the Finish block. If a block
cannot be reached from the Start block, then the instructions in that block will never execute. If the path from a
block does not eventually reach the Finish block, then the
logic will never properly terminate. A common error is to
terminate a path at a result or action, other than Finish.

8 Finite Difference Methods and MATLAB

502
Enter Outer Loop
of the Nested Loops

Instructions for
actions or computations
Instructions for
actions or computations
Enter Inner Loop

Instructions for
actions or computations

Instructions for
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
actions or computations

Instructions for
actions or computations

No

Outer
Loop Tasks
Done ?

Instructions for
actions or computations

No

This is a logical dead-end

Conditonal
True ?

Yes

Yes

Instructions for
actions or computations

Exit Outer Loop

Fig. A8.6 Nested loops

2. Have only one path into a block. If there are multiple


paths leading to the block, have them join at a node before
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
blocks to read from bottom to top.
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
through a set of instructions. In both flowcharts, a counter
variable, n, is assigned the value of one beforethe logic enters the loop. The loops in both flowcharts contain identical
instructions. The first instruction in both loops computes the
output variable, y, by squaring the counter variable, n. However, the order of the two remaining blocks is different in

Finish

Fig. A8.8 Flowchart with dead-end logic

the two flowcharts. In Flowchart A, the counter variable, n,


is incremented, or increased, by one with the instruction:
n = n +1
before the logic enters the decision block with the relational expression, n==10? In Flowchart B, the logic enters
the decision block and evaluates the relational expression,
n==10?, before the counter is incremented.

MATLAB

503

Fig. A8.9 a Two branches entering one block, logical but can
be difficult to read. b Preferred
flowchart. Flow merges at a
node before entering the block

Instructions for
actions or computations

Instructions for
actions or computations

Instructions for
actions or computations

Instructions for
actions or computations

Instructions for
actions or computations

Instructions for
actions or computations

Start

Instructions for
actions or computations

Instructions for
actions or computations

Instructions for
actions or computations

Initialize counter variable


n=1

Instructions for
actions or computations

No

Compute output variable y


y = n2
Done ?

Increment counter by one


n=n+1

Yes

Fig. A8.10 Flowchart with upside down flow: logical, but difficult to
read and, therefore, error prone

The program shown in Flowchart B will loop 10 times,


as n is incremented from n = 1 to n = 9. When n = 9, the
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.
The effect of the relative position, where the counter variable is incremented within the loop, and the relational expression, 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.

MATLAB
Modern programming languages provide an environment,
which is a desktop and a set of windows. The menu, Desktop, contains items which allow you to open needed windows. We will use the Command window and the Editor
window. MATLAB retains the configuration of the environment when the program is closed, and returns to that con-

No

n == 10 ?

Yes

Finish

Fig. A8.11 Flowchart A

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
previous user.

MATLAB Environment
The MATLAB environment includes the Command window, the Editor window, and the Help window, among others.
The Editor is well hidden. MATLABs Editor opens, when

8 Finite Difference Methods and MATLAB

504

Start

Initialize counter variable


n=1

Compute output variable y


y = n2

Increment counter by one


n=n+1

No

n == 10 ?

Yes

Finish

Fig. A8.12 Flowchart B

the user clicks on the New Script button. A script is a short


program. The term, script, came into use after the size of
programs became enormous. Scripts, or programs, are written in the Editor window, saved as an m-File (a file with
extension file_name.m), and then run from the Command
window. MATLAB does not permit spaces in file names.
MATLAB looks in a number of locations to find m.files,
including the default directory. The collection of directories
and folders checked for m.files is known as MATLABs
path. If you write a script and attempt to save it to your flash
drive, you will see a MATLAB message stating the folder is
not in MATLABs path, and ask if you would like to add it.
Accept this choice, so that MATLAB will look at your flash
drive for a script.
A MATLAB program is run, by typing the name of the
m-file without the extension, .m, at the prompt in the
command window. Example: If you have an m-file named
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
you open MATLAB on a public computer, click on the Set
Path button in the Environment tab, and add your flash drive
folder to the path. If you forget to do this, and you have chosen a common name for your script, then MATLAB may find
someone elses script of the same name saved to one of its
default directories, and run that one instead. It may take a

while to recognize that script is not yours, particularly if you


failed to add comments.
MATLAB is an interpreter, meaning, it will happily execute a script, or the same commands entered at the prompt
in the Command window. The variables used during a work
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.
However, there is a dark side to MATLABs workspace.
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
simply forget, and omit that line of code. If you run the script,
and tau was not used in any previous calculations, so that it
is not in the workspace, MATLAB will identify the error. A
ding sounds. The Command window contains an error message, indicating that an undefined variable was used in line x
of the script. On the other hand, if you omit the definition of
tau in the script, but another tau already exists in the workspace, MATLAB will use the workspace value. The script
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
of an existing variable. So if a MATLAB execution error occurs, or the results of a successful execution of a script dont
make sense, clear the workspace, and run the script again.
The entire workspace is cleared by selecting the title bar of
the Workspace window, right clicking, and choosing Clear
Workspace. Individual variables are deleted, by selecting the
variable in the Workspace window, right clicking, and choosing delete. MATLABs retention of variables is a particular
nuisance, when debugging a script. It is common to change
the number of iterations within a for loop, while debugging.
When results of a computation are written to a vector, or an
array, and the number of iterations is increased, no problem.
If the number of iterations is decreased, the values beyond
the most recently overwritten elements are unchanged. A
plot may make sense, nearly to the end, before becoming
quite strange. When faced with inexplicable results, clear
MATLABs workspace first, before tearing apart your script.

Variables
MATLAB does not require variables to be declared, meaning the types of data assigned to variables do not have to
be identified before they are used. Integer and floating point
variables are identified by their use. Integer variables are assigned integer values, i.e., numbers without a decimal point.

MATLAB

505

Floating point variables are assigned either decimal numbers


or numbers expressed in scientific notation.
All variables must have a name. In MATLAB, variables
names must begin with a letter and may contain numbers
and underscores, but no other special characters. MATLAB
is case sensitive.
Values are assigned to variables with an assignment statement. 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.

Many computer languages will not allow you to use a logical variable in a numerical calculation. MATLAB is more
forgiving. Type

PDVV 
0$7/$%HFKRHV\RXULQSXWDV

MATLAB was written to facilitate vector-matrix mathematics. MATLAB does not require the user to declare a variable
as a scalar, vector, matrix, or an array. Scalar, vectors, and
matrices have their conventional mathematical definitions,
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
columns. MATLAB distinguishes between matrices and arrays by the way one can operate on them. Matrix operators
obey the rules of linear algebra. Array operators operate element by element.
The following assignment statement creates a vector or an
array. Important: Note the use of square brackets!

!!PDVV 
PDVV 

An assignment statement assigns the value of the right side
to the variable named on the left side. Assignment statements
are not equations! Typing the statement in the reverse order
(capitalizing the M in Mass to make it a different variable)
produces an error.

 0DVV
MATLAB thought you tried to name a variable, 3.5. This
violates two of its rules: (1) variable names must start with a
letter, and (2) the only permissible non-alphanumeric character is an underscore. Periods, points, or dots are not permitted in variable names.
Variables can be assigned strings of characters. Characters or strings of characters are identified by pairs of single quotes. Example: type the following valid assignment
statement

$ (QWURS\URFNV
Note how MATLAB color-codes the input as you are typing.
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 either true or false to the variable. MATLAB recognizes true
as equivalent to the Boolean value 1, and false as equivalent
to the Boolean value 0. When you type

D WUXH

& DPDVV
and you will find the result is 4.5000.

Scalar, Matrix, and Array Variables

9 >@
Note spaces separate the elements within brackets. Alternatively, the elements within brackets can be separated by commas, or by a mixture of commas and spaces.

9 >@RU9 >@


A semicolon terminates a row. Example:

0 >@
Is echoed to the screen as:

!!0 


Individual elements can be addressed using row and column
subscripts within parentheses. Example: The element, 7, is in
the second row and third column of the matrix (or array), M4.

!!9  
DQV 


MATLAB echoes

D 

The lowest index in MATLAB is 1, not 0!

8 Finite Difference Methods and MATLAB

506

MATLAB is very tolerant with vector-matrix mathematics. It will execute operations which are not defined in linear
algebra, by interpreting your vectors and matrices as arrays.
For example, in conventional vector-matrix mathematics, if
b is a scalar, and B is a row vector, e.g.,

E 
and

% >@
then the sum

& E%
is not defined. MATLAB, however, will execute this statement and echo

Table A8.6 MATLABs arithmetic operators


Operator

Description

Addition

Subtraction

.*

Array multiplication

./

Array right division

.\

Array left division

Unary plus

Unary minus
Colon operator

.^

Array power

Transpose

Complex conjugate transpose

Matrix multiplication

Matrix right division

Matrix left division

Matrix power

& 

MATLAB has a type of array, called a cell array, to handle
non-numerical data. It refers to the element of an array as a
cell. The assignment statement for an array uses braces
(which MATLAB refers to redundantly as curly braces).
The assignment statement

*BFHOOBDUUD\ ^WZRWKUHHIRXU`
is echoed as

*BFHOOBDUUD\ 

WZR

WKUHH

IRXU

Note the individual cell data are maintained. Contrast this
result with a conventional array, defined by the use of square
brackets, containing the same data. The assignment statement

*BDUUD\ >
WZR

WKUHH

IRXU
@
is echoed as

*BFHOO 
WZRWKUHHIRXU

Operators

All programming languages provide at least three different types of operators: arithmetic, relational, and logical.
Arithmetic operators are used in computations. Relational
operators are used to define conditions to control the flow
of a programs execution. Logical operators are used for
Boolean logic.

Arithmetic Operators
Operators and the symbols used for operators vary tremendously across programming languages. The only operators,
whose meaning will not waver, are the arithmetic operators
for multiplication, division, addition, subtraction, as long as
they are used as binary operators (with two operands), as
they were in elementary school. Many operators, other than
dear Aunt Sally, are included in a programming language.
The set of operators varies with the intended application of
the language. For example, languages developed to program
webpage applications will have operators to parse through
strings of character data. MATLAB was written for scientific
and engineering computation, so it has more mathematical
operators than most languages.
Symbols to represent operations were limited to those on
the computer keyboard. Consequently, operators are represented by combinations of symbols that made sense to the
programmer, but do not follow any convention and vary
widely between languages. For example, exponentiation of
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.
We will use these arithmetic operators in MATLAB: the
elementary school arithmetic operators of multiplication,
division, addition, subtraction; the unary plus and minus
operators (which are simply the plus and minus signs); and
exponentiation. The complete set of arithmetic operators is
presented in Table A8.6.
This set requires some explanation. MATLAB uses the
terms, matrix operations and array operations, to distinguish between conventional linear algebra operations performed with vectors and matrices from those operations not
defined in linear algebra, which are performed with two arrays of the same size and shape, or on an array and a scalar.

MATLAB

507

An array operation is performed element by element,


which means the operation is performed repeatedly, using
elements in the same row and column position in each of
the two arrays, until the operation has been performed on
all the pairs. Note array operations are preceded with a dot
(.), except for addition and subtraction, which are the same
as matrix addition and subtraction, since they are performed
element by element. For illustration, define two square arrays A and B:
a11 a12
A=
a21 a22

b11 b12
and B =
b21 b22

Array Multiplication:
a11b11 a12b12
A. * B =
a21b21 a22b22

Table A8.7 MATLABs array and matrix operators


Operator Description
Conventional Linear
Algebra?
+
Addition
Yes

Subtraction
Yes
.*
Array Multiplication
No
./
Array Right Division
No
.\
Array Left Division
No
+
Unary Plus
Yes

Unary Minus
Yes
Colon Operator
No
.^
Array Power
No
.
Transpose
Yes

Complex Conjugate Transpose


Yes
*
Matrix Multiplication
Yes
/
Matrix Right Division
No
\
Matrix Left Division
No
^
Matrix Power
Yes

then the C transpose is

Array Right Division:


a11
b
11
A. / B =
a
21
b
21

a12
b12

a22
b22

c11 c 21
C. = c12 c 22

c12 c 22
MATLAB recognizes both i and j as imaginary numbers
when a complex number is entered as
or

z = 2 + 3i

Array Left Division:


b11
a
11
A. \ B =
b
21
a
21

b12
a12

b22
a22

z = 2 + 3j

However, when MATLAB echoes back a complex number, i


is used as the imaginary number.
The MATLAB operator, which transposes a complex
number to obtain its complex conjugate, is .

]
  L
Array Power:
a b11
A. ^B = 11b21
a
21

a12b12
a22b22

An exception to the syntax of preceding array operators with


a dot is the transpose operator, . . The transpose operator
can operate on vectors, matrices, and arrays. It is a unary
operator, since it has only one operand. The transpose operator interchanges the order of the elements to move an element from the rth row and cth column to the cth row and rth
column. Example: If
c11 c12
C=
c 21 c 22

c12
c 22

The Colon operator, :, creates a range variable. Example:


Type:

!!D 
and MATLAB will echo

D 

In this case, the variable, a, is now a vector or an array. The
fifth element is

!!D  
DQV 


8 Finite Difference Methods and MATLAB

508

To increment by a value other than unity, indicate the first


value of the series, the increment, and the upper limit. Example: The statement

!!E 
yields two values: the initial value and one increment. The
second increment would exceed the upper limit of the series

E 

MATLAB includes two odd operators, Matrix Right Division and Matrix Left Division. Matrix division is NOT defined in linear algebra. The matrix division implemented
in MATLAB is a shortcut for matrix inversion. Although matrix division may be advantageous in some circumstances,
it is better to use matrix inversion.

Relational Operators
A computer languages arithmetic operators are used for
crunching numbers. The languages relational operators are
used to compare the values of variables. The operators allow
a computer to think, by branching the flow of the execution on the outcome of relational expressions. Relational operators produce a result that is a Boolean, or logical, value,
i.e., either true or false. For example, if A and B are numbers,
then the answer to the question, Is A greater than B? must
be either true or false.
MATLABs relational operators are given in TableA8.8.
Table A8.8 MATLABs relational operators
<
>
<=
>=
==
~=

Less than
Greater than
Less than or equal to
Greater than or Equal to
Equal
Not equal

Only the last two operators require discussion. Note the


relational operator, Equal, consists of two equal signs. A common error is to confuse the single equal sign, which is the assignment operator, and the double equal signs, which signify a
relational equal operator. Also note the symbol for Not Equal,
which is unusual. Most programming languages use the exclamation point, ! for the Not, or logical inversion operator.
Most engineers would read ~= as approximately equal, which
is why MATLAB chose it to represent Not Equal.
The distinction between approximately equal and equal is
a warning for the use of the Equal operator in an if statement.
The Equal relational operator requires an exact mathematical
or logical equality, not an engineering equality. An engineer

would accept the statement, 0.99999==1, as true in most


circumstances, but to MATLAB, the statement is false. Do
not use the Equal relationship operator with floating point
(scientific notation) variables, since they are likely to fail the
comparison due to minute numerical error.

Logical Variables
Boolean logic is performed in MATLAB using logical
variables which are defined or created by assigning them a
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 positive, the result is true. If the argument is negative or zero, the
result is false.
Logical Operators
MATLAB provides the three fundamental Boolean operations, And, Or, and Not, in two forms, as operators and as
functions.
Table A8.9 MATLABs logical operators
Logical operation

Equivalent function

A&B
A|B
~A

and(A, B)
or(A, B)
not(A)

Boolean operators are used with relational operators to


construct conditional statements that control which program
instructions are executed. The Boolean operators are best
described using truth tables, where A and B are Boolean
variables which must have values of either true or false.
Table A8.10 and truth table
A

A&B

Table A8.11 or truth table


A

A|B

Table A8.12 not truth table


A

~A

MATLAB

509

Programming Statement Syntax


A computer programming languages syntax is why computer code is called code. A programming languages syntax
is restrictive and specific, so that the characters read by the
machine can be interpreted as instructions.

Assignment Statements
A program is a list of instructions coded in the programming
languages syntax. It is important to recognize that program
statements are not equations. For example, the following assignment statement in the flowcharts Figs.A8.12 and A8.13
is a valid statement in all computer languages, even though it
is not a valid equation, except in Boolean algebra
Q Q 

is executed. Echoing the code and the results to the monitor


during execution of a program slows the execution considerably. Consequently, you will want to place semicolons at the
end of line in loops which will execute repeatedly.

Control Flow Statements


if Statement
A fundamental instruction in all languages is an if statement. The syntax varies with the language but the essence is
if (expression A Relational Operator expression B) is
true then execute the following instructions
In MATLAB, the syntax of an if statement is:

if (expression A Relational Operator expression B)


Instruction executed if conditional is true
Instruction executed if conditional is true

Instruction executed if conditional is true


end

As an algebraic equation, this expression is nonsense. No


number can be equal to itself plus one. As a program statement (or instruction), however, the expression not only
makes sense, it is very important. It is the instruction to increment the value of the variable, n, by one. Recall this expression is called an assignment statement, and a variable
name is an address in the computers memory. This assignment statement is understood by the computer as saying,
Write (or copy) the contents of the memory address(es) assigned to the variable, n, to the specialized hardware in the
central processing unit called the accumulator, which performs operations on binary numbers. Add one to the value
in the accumulator. Write the contents the accumulator to the
memory address(es) assigned to the variable n.
Assignment statements are read from right to left. The
right side can be any valid mathematical statement expressed
using MATLABs notation, with the restriction that variables
cannot be used, before they have been assigned a value. If
we use the expression, n = n + 1, before giving n an initial
value (referred to as initializing the variable), a red error
message would state
??? Undefined function or variable n

The following code would execute properly


Q Q 

if (expression A Relational Operator expression B)


Instruction executed if conditional is true
Instruction executed if conditional is true

Instruction executed if conditional is true


else
Instruction executed if conditional is false
Instruction executed if conditional is false

Instruction executed if conditional is false


end

Many languages require a character, such as a semicolon,


to indicate the end of an instruction. MATLAB does not.
In MATLAB, a semicolon at the end of a line prevents the
code from being written to the monitor, when the program

else Statement
The else instruction is used to divide the block of instructions between an if statement and its corresponding end
statement into two blocks. The first block is executed, if the
conditional statement is true. The second block is executed,
if the conditional statement is false. Example:

Q 

The instructions appear between the conditional statement


with the relational operator and the end statement. Important:
MATLAB is case sensitive. The keywords, if and end,
are both lower case. MATLAB highlights the leading if
and the closing end in blue to demarcate the limits of the
instructions, which will execute if the conditional statement
is true. To further demarcate the if statement as a block of
code, it is good practice to indent the instructions, which are
executed by the if statement.
MATLAB and many other languages provide two additional instructions to supplement the if statement. They are
else and elseif.

8 Finite Difference Methods and MATLAB

510

elseif Statement
The elseif instruction is a second conditional statement
that follows an if statement. It is executed only when the
conditional statement of the if statement is false. Example:

if (expression A Relational Operator expression B)


Instruction executed when if conditional is true
Instruction executed when if conditional is true

Instruction executed when if conditional is true


elseif (expression C Relational Operator expression D)
Instruction executed when elseif conditional is true
Instruction executed when elseif conditional is true

Instruction executed when elseif conditional is true


end

while Loop
A while loop uses a conditional statement to control the repeated execution of a block of code. The structure of a while
loop resembles an if statement.
while (expression A Relational Operator expression B)
Instructions
Instructions
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
while statement is executed repeatedly, as long as the conditional statement remains true. It is possible to code an infinite loop. If you ever need to interrupt the execution of a
MATLAB program, type CTRL+C simultaneously.
The following code with a while loop executes the computation flow charted above.
n=1
while (n ~= 10)
y = n^2
n=n+1
end

for Loop
A for loop is a block of code which executes for a fixed
number of times. The number of executions is controlled by
giving a counter variable a range, using the colon operator.
The block of code which executes repeatedly is demarcated
with the key word, end. A MATLAB for loop has the following form:

for counter = lower integer limit: upper integer limit


Instructions
Instructions
Instructions
end
The following code would execute the computation flow
charted above
for n = 1:10
y = n^2
end
Nested for loops are often used. The following code produces a matrix, or array, named A with two rows and three
columns. The inner for loop increments through the columns. The outer for loop increments through the rows.
for r = 12
for c = 13
A(r,c) = (r+c)^2
end
end
The result is

$ 



Example: for Loop and while Loop


The following example is a first-order finite difference algorithm, programmed using a for loop to control the number of
iterations through the loop:
1 
WBHQG 
D 
E 
[   
X 
7 WBHQG1
W   
IRUQ 1
'[ Q  D [ Q 7E X 7
[ Q  '[ Q [ Q 
W Q  Q 7
HQG
SORW W[

MATLAB

Here is an alternative program, using a while loop to control


the number of iterations

1 
WBHQG 
D 
E 
[   
X 
7 WBHQG1
W   
ZKLOH 1 !Q
'[ Q  D [ Q 7E X 7
[ Q  '[ Q [ Q 
W Q  Q 7
Q Q
HQG
SORW W[

Comments
It is essential to include comments in a computer program
to describe the algorithm and identify variables. Comments
serve to document the program and make it easier to read.
Comments must be distinguished from the programs code.
It seems every language uses a different symbol to identify
a comment. MATLAB uses the percent sign, %. MATLAB
will ignore all characters which follow a % on the same line.
Comments may be included on the same line as a program
instruction, or on a separate line. Example:
% This is a comment.
A = B + C % This is a comment following code.
Compare readability of the m-file code for the spring-massdamper system above, with the m-file with comments below.
Commented script for the spring-mass-damper system
state equations.
% Second order system state equation solver
%
N = 2000
% number of steps
t_end = 100.0 % duration of simulation
%
%
Parameter values
M = 3.0
K = 1.0
B = 0.2
%
Coefficient matrix elements
a11 = -B/M
a12 = -1/M
a21 = K
a22 = 0.0

511

%
Input matrix elements
b1 = 1/M
b2 = 0.0
%
Initial values of state variables
%
x1(1) = 0.0 % mass velocity v1g
x2(1) = 0.0 % Force acting through spring FK
%
u = 10.0
% step input F
T = t_end/N % time step
t(1) = 0.0
% initial element of time vector
%
%
Difference equation loop
for n = 1:N;
Dx1 = (a11*x1(n) + a12*x2(n) + b1*u)*T;
Dx2 = (a21*x1(n) + a22*x2(n) + b2*u)*T;
x1(n+1)= Dx1 + x1(n);
x2(n+1)= Dx2 + x2(n);
t(n+1) = n*T;
end;
%
plot(t,x1,t,x2)

plot Statement
The MATLAB command to create an xy plot is

SORW [\
where x and y are column vectors, or one-dimensional arrays
of equal length. Multiple plots can be created using the same
axes, by adding x, y pairs

SORW [\[\
If you wish to plot two output variables, y1 and y2, against
the same independent variable, t, then you must repeat the
independent variable in each x, y pair

SORW W\W\

Plotting and Labeling Multiple Figures


MATLAB plots onto a graphic object called a figure.
In other words, a plot is the trace drawn on (or in) a figure.
MATLABs function, plot (x, y), draws a plot of the x, y pairs
of the vector x and the vector y on a figure. If a figure has not
been created, then the function, plot (x, y), creates a figure.
If a figure exists, then plot (x,y) wipes that figure clean, and
draws a plot on it.
Plotting multiple figures requires creating a new figure
with a number as its handle to identify it. The command is
ILJXUH KDQGOHQXPEHU 

512

The following MATLAB code yields only the last plot, the
plot of x3 vs. y3, because the figure created by the first plot
would be reused twice.

SORW [\
SORW [\
SORW [\
The command figure creates a graphic object. The following
MATLAB code would yield three new plots numbered successively from 1, each time the code was run. Run the code
four times, and a total of 12 plots would be created and stay
open.

ILJXUH
SORW [\
ILJXUH
SORW [\
ILJXUH
SORW [\
There are two methods to limit the number of figures created
by successive runs of the same code. One method is to use
the close command, prior to the figure command. The command, close all, will close all open figures.

FORVHDOO
ILJXUH
SORW [\
ILJXUH
SORW [\
ILJXUH
SORW [\
The second method is to identify the figure with a handle,
which, in our case, is a number. This is done by adding the figure number to the command, as an argument within parentheses. The following code creates the plot of x1 vs. y1 on figure
number1. If figure number 1 does not exist, the command, plot
(x1,y1), creates it. If figure number 1 does exist, then the command, plot (x1,y1), will wipe it clean and reuse it.

ILJXUH  
SORW [\
ILJXUH  
SORW [\
ILJXUH  
SORW [\

Labeling Plots
If you are going to plot multiple figures, you must title them
to identify them. A plots axes should be labeled to identify
the variables and units.

8 Finite Difference Methods and MATLAB

Figure titles are added with the function title(Plot Title).


X-axis labels are added with the function, xlabel(meaningful
x-axis label). Y-axis labels are added with the command,
ylabel(meaningful y-axis label). The text used in these
functions must be enclosed in single quotes. Example:

WLWOH
3UREOHPH6SXU*HDU6\VWHP

[ODEHO
WLPHVHFRQGV

\ODEHO
6KDIW7RUTXHQHZWRQV

Formatting Plots
There are various symbols, line types, and line colors available in MATLAB. They are selected by adding arguments
to the plot(x, y) command after the pair of vectors. See
TableA8.13. The default is a solid, black, thin (0.5) line. Formatting the line type, color, and width is particularly helpful
when there are multiple plots (traces) on a figure. Line type
formats are coded graphically; solid is '_', dashed is '', dotted is ':', and dot-dash is '.'. Black is k, blue is b, red is
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
quotes around b and LineWidth.
SORW [\


E

/LQH:LGWK

Color is assigned with a color letter code in single quotes.
The colors are cyan, magenta, yellow, black, red, green,
blue, and white. The line width is specified by LineWidth
in single quotes and without a space between the words, followed by a comma and a number indicating the width before
the plot command. Examples:
figure(1)
plot(rkt,rkx1,'k','LineWidth',2)
title('Problem 8.5.e, Worm Gear System')
xlabel('time, seconds')
ylabel('Angular Velocity of Gear, rad/sec')
%
figure(2)
plot(rkt,rkx2,':','b','LineWidth',2)
title('Problem 8.5.e, Worm Gear System')
xlabel('time, seconds')
ylabel('Angular Velocity of Load, rad/sec')
%
figure(3)
plot(rkt,rkx3,'-.','r','LineWidth',2)
title('Problem 8.5.e, Worm Gear System')
xlabel('time, seconds')
ylabel('Shaft Torque, newtons')

MATLAB
Table A8.13 MATLABs line type, color, and marker format codes
Line-type format Code Line marker format code
Solid line (default) +
Plus sign
-- Dashed line
o
Circle
:
Dotted line
*
Asterisk
. Dash-dot line
.
Point
Line Color Format
x
Cross
Code
r Red
square or s
Square
g Green
diamond or d Diamond
b Blue
^
Upward-pointing triangle
c Cyan
V
Downward-pointing
triangle
m Magenta
>
Right-pointing triangle
y Yellow
<
Left-pointing triangle
k Black
pentagram
Five-pointed star
or p
(pentagram)
w White
hexagram
Six-pointed star
or h
(hexagram)

A Brief History of Control Characters


The syntax MATLAB uses to format data is a mixture of
format strings from FORTRAN, a programming language
which dominated engineering applications in the 1960s and
1970s, and control characters, which are a century older,
dating from 1870s International Telegraph Alphabet, the
5-bit code used to drive the first teletypes, invented by Emile
Baudot. The unit baud, the SI unit for symbols per second, is
named for Baudot.
The carriage return control characters, /r, refer to the carriage of a typewriter. The first teletypes were electromechanical typewriters. The carriage of a typewriter is the assembly,
which carries the paper; the platen, a rubber-covered roller
which backed up the paper; the feed rollers; and guides. The
carriage moved the paper from right to left in front of the key
bars. The linkages connected the key of the keyboard with the
type slug, which struck the ink-saturated cloth ribbon, and
transferred the image of a character to the paper. Teletypes
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 minicomputers. Teletypes were used in place of computer monitors,
because the teletypes were less expensive, even though they
cost approximately $700 in 1970. To put that in perspective,
a Volkswagen cost less than $2,000 in 1970. Teletypes died
out, when the introduction of inexpensive dot matrix printers finally made them obsolete in the late 1970s. You will
never work with one, but you will see them in movie scenes
of newspaper newsrooms from the twentieth century. (Soon,
you will only see newspapers in movies too.)

Programming a Function in MATLAB


See Sect.8.3 of this chapter.

513

Reading From and Writing To Files


We will need to read data from files and write data to files,
in order to use MATLAB with laboratory data. MATLAB
conveniently has interactive methods using a graphical user
interface. However, our present focus is on programming.
Because Microsofts Excel is so widely available, it is common for engineering software, including MATLAB and
Mathcad, to be able to read and write numerical data to and
from an Excel worksheet. MATLAB also includes more
general and flexible functions for reading and writing data
to files, which we would need to use for reading and writing strings of characters or data to a text. Unfortunately,
complex syntax is the price of flexibility. The programmer
is responsible for opening and closing files and formatting the data. The latter is a throwback to the days of
programming computers using punch cards. We will stick
with Excel worksheets.
In order to use MATLABs syntax to read and write Microsoft Excel files, we need to either assemble an array to
write to an Excel worksheet or disassemble an array read
from an Excel worksheet.

Concatenation of Arrays
Concatenation is the operation of joining arrays to create
a new array. It is easiest to understand the operation through
illustration. Starting with arrays A and B, where
A=

a11 a12
a21 a22

and

B=

b11 b12
b21 b22

Arrays A and B can be concatenated as either


a11
a21
C=
b11
b21

a12
a22
b12
b22

or

D=

a11 a12
a21 a22

b11 b12
b21 b22

MATLABs concatenation function, cat( , , ), takes three


arguments. The first argument, , controls whether arrays
and are concatenated with above and below, or side by
side with to the left and to the right. The argument, , is
the dimension along which the concatenation occurs. We
will most often work with one or two-dimensional arrays.
A one-dimensional array has only rows, xrow. Therefore, the
dimension of the element index for rows is 1. A two-dimensional array has rows and columns, yrow,column. The dimension
of the index for columns is two. The MATLAB syntax which
produced arrays C and D is:
C = cat (1,A,B)

and

D = cat ( 2,A,B )

8 Finite Difference Methods and MATLAB

514

If you prefer not to work with the dimensions, MATLAB


also provides the functions, vertcat() and horzcat()
C = vertcat ( A,B)

and

D = horzcat ( A,B )

The most common error in concatenating arrays is to attempt


to concatenate two-dimensional arrays which do not fit
next to each other properly. Two arrays must have the same
number of columns if they are to be concatenated above and
below, or the same number of rows if they are to be concatenated side by side.
By the way, when MATLAB help refers to multidimensional arrays, it means arrays with more than two dimensions.

Extracting a Row or Column from a


Two-Dimensional Array
MATLAB uses the colon operator : as shorthand for a
range variable which indexes by unity. The lower limit, j,
and the upper limit, k, can be assigned as j:k.
When a range variable is used as one of the indexes of
an array, and no limits are assigned to the range variable, in
other words, the colon is used by itself, then the range variable is assigned the limits of the row or column of the array.
This shorthand notation makes it simple to extract a single
column or row from an array. For example, the following
code extracts the second column from the array C and assigns it to the variable, data2:
data2=C(:,2)
yielding the column vector
a12
a22
data2 =
b12
b22
We achieve the same result using the lower limit of one and
the upper limit of four

GDWD &  

Reading and Writing to a Microsoft Excel


Worksheet File
The MATLAB syntax for reading a Microsoft Excel worksheet file and assigning the data to the variable, data3, is
GDWD [OVUHDG FRPSOHWHILOHSDWK 
Note the name of the file to be read is the complete file path,
enclosed in single quotes. If you do not see the file path in
the colored address bar at the top of the dialog, use the

Tools menu to change the file settings. The complete file


path will have a form similar to
C\Documents and Settings\Class User\My Documents\
Systems\datafile.xls
It is easier to read from or write to a flash drive, which will
have a short full path name. If your flash drive were assigned
the drive letter R, then the full path name would be

5?6\VWHPV?GDWDILOH[OV
The syntax for creating Excel worksheet file and writing to
it is

[OVZULWH FRPSOHWHILOHSDWKDUUD\WREHZULWWHQ 
where array to be written is the variable name of the array.
Example: Say you plot the result of a RungeKutta simulation, and wish to save it as a Microsoft Excel worksheet
file. The plot() function requires vectors holding the x data,
time in our case, and y data, say force F. We will concatenate these data into a two-dimensional array, name the array,
plot_output, and write the data as an Excel worksheet file to
a flash drive R with the following code.

SORW W) 
SORWRXWSXW KRU]FDW W) 
[OVZULWH
5?0(?UHVXOWV[OV
SORWBRXWSXW 

Reading and Writing to a Text File


Although we refer to many different types of files, such
as a Microsoft doc file, a Mathcad mcd file, a MATLAB m file, etc., what we are referring to are the filename extensions (what follows the final dot in the file
name), which allow a program to recognize a family of files
it can read and write to. There are, in fact, only two fundamental file types binary files, in which data are encoded
in binary and text files, in which data are encoded in characters. Most of the files created by the programs we work
with are text files.
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 independence, in that you do things the way MATLAB wants you to.
Reading and writing text files in MATLAB is more involved
than Excel files. You regain your freedom, and the responsibility that comes with it, to take care of the details.
There are three steps for writing to or reading from a text
file.
1. Open the file, meaning
a. identify or create a file
b. declare whether you plan to write to it or read from it
c. assign the file to a variable name

MATLAB
Fig. A8.13 fprintf function, annotated
to show the association of format strings
and data

515

fprintf(File_Variable_Name,'%format %format new line',data 1,data 2);

2. Write (or read) line by line using a loop


a. format the data by defining the width of the field
and the data type
b. end the line of data with a carriage return or new line
control character
3. Close the file
The command to open a file is

format, which is a conventional decimal number, or e to


mean exponential, or scientific notation. Note that f must
be lower case, but either e or E is acceptable. Example: The
number, 31,415.92, converted using both fixed and exponential format strings:
I


)LOHB9DULDEOHB1DPH IRSHQ ILOHBQDPHZ 


Note the use of SINGLE quotes around the file_name and
the argument w, where w means the file is opened to
write to. Example:

ILG IRSHQ
/DEB5.B/LQBW[W

Z

The command to write a line of two formatted data to an
open file is
fprintf(File_Variable_Name,'format format new line',
data 1,data 2);
The command fprintf() stands for File Print Formatted.
The syntax MATLAB uses to format data is literally from
the 1960s. All of the formatting information is between two
single quotes. A format string begins with a percent sign,
%, to indicate what follows is the format code. There must
be as many format strings as there are data. The first format
string formats the first datum. The second format string formats the second datum. Note there are no commas between
format strings, just a space, while the data variables are separated by commas.
An example is

ISULQWI ILG
II?Q
UNW NN UN[ NN 
The first number following the percent sign is the maximum number of numbers to be written. This is the field
width. It will be padded with blanks to the left of the most
significant digit. Next, there is a decimal point followed by
a number. The number after the decimal point, called the
precision, is the number of digits to be written after the
decimal point. The format string ends with a letter called
a conversion character, which specifies the notation the
number is to written in. We will use either f to mean fixed


(



(

A common error is to use too narrow a field width. Remember 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.
The formatting information ends with control characters which date from the 1930s and teletypes. Two useful
control characters, when written to a file, are \n, which commands a new line and \r, which commands a carriage return.
Closing the file is the easy part. The command is

IFORVH )LOHB9DULDEOHB1DPH 
Forgetting to close a file means it cannot be opened and read.
If you can access the file you created, check to see if you
closed the file in your code.
The following is MATLAB code which writes the arrays,
rkt and rkx1, to the file, Lab3_RK_Lin_09172014.txt.
fid = fopen('Lab3_RK_Lin_09172014.txt','w');
for kk=1:N;
fprintf(fid,'%8.4f %12.8f \n',rkt(kk),rkx1(kk));
end
fclose(fid);

MATLABs step() and impulse() Functions


MATLABs Control System Toolbox contains the functions, step() and impulse(), which calculate the unit step
response and the unit impulse response of a transfer function. The ability to calculate a time-domain response from
a transfer function is a tremendous convenience. The step()
and impulse() functions take the same arguments and work

516

the same way. The following discusses the function step()


but also applies to impulse(). If step() is given the name of
a transfer function, say Model, as its single argument, tend =
0.5, step(Model), the duration of the response may or may
not meet our needs. We can set an end time for the calculation, by adding a scalar argument, say step(Model, tend),
to force unit step calculation to run from t = 0 to t = tend .
We can also pass step() a vector with the times at which to
evaluate the unit step response. In our application, the time
vector would be that of the observed response, step(Model,
timeVector).
The following MATLAB m-file program reads the RCL
circuit discharge data used in the Mathcad code above from
Excel files, plots the data, then calculates and plots the impulse response of the model.
% Program Lab 0 RLC Discharge Data and Model
%
% Read the time and voltage data from Excel files.
%
time = xlsread('j\timeVector090214.xls');
data = xlsread('j\dataVector090214.xls');
%
% Close all existing figures.
%
close all;
%
% Create a figure for the plot.
%
figure;
%
% Plot the x, y vector pair. The format string is '.b'.
% No line type is indicated. The period indicates
% that the "marker", or data point type, is a point.
% Normally, the default when the line type is omitted
% is a solid line. However, when no line type is
% indicated and a line marker is given, the plot will
% have no line, just the line markers.
% The "b" stands for blue.
%
plot(time,data, '.b');
% Assign values for the natural frequency,
% damping ratio and the gain of the model.
%
wn = 6.293E5
z = 0.258
K = 3.29E-6
%
% Define "s" to be a transfer function in order to use
% it as the Laplace variable.
%
s = tf('s')
%
% Assign the transfer function to the variable "model".

8 Finite Difference Methods and MATLAB

model = K*wn^2/(s^2+2*z*wn*s+wn^2);
%
% Retain the previous plot with the next.
%
hold on
% Calculate the unit impulse of the model at the
% times in the vector time. Assign the model
% response to the variable modelResponse.
% Assigning the output of impulse( ) to a variable is
% necessary to format the plot in code. This is also
% true for the functions step( ) and rlocus( ).
%
modelResponse = impulse(model,time);
%
% The formatting string is '-k'. "-" means a solid line.
% "k" stands for black. The remaining attributes
% are set as pairs of 'attribute name', attribute value.
% Here 'LineWidth' is set to 2.
%
plot(time,modelResponse,'-k','LineWidth',2);
%
% The axes labels and the plot title are assigned
% after the plot is created.
%
xlabel('time, seconds');
ylabel('volts, VDC');
title('RLC Circuit Data vs. Model');

Vector Calculations in MATLAB


MATLAB has two functions to determine the number of elements in a vector, length( ), used on vectors, and size( ),
which is used on arrays of any number of dimensions, including vectors. When size( ) is used on a vector, it will return the number of rows, n, and the number of columns, 1, as
a row vector, [n 1]. Remember that a space or a comma can
be used to separate elements in a row in MATLAB. Rows of
vectors and arrays are separated by semicolons.
Calculation of the integrals of the error, the absolute value
of the error, and square of the error is performed in MATLAB
with for end loop. The MATLAB m-file program follows:
% Program to calculate of the integral of the error,
% the absolute value of the error and the error
% squared between a models response and the
% observed response.
%
% Read Excel files with the time and amplitude data.
%
time = xlsread('j\timeVector090214.xls');
data = xlsread('j\dataVector090214.xls');
%
% Assign values for the natural frequency,
% damping ratio and the gain
%

References and Suggested Reading

517

RLC Circuit Data vs. Model

1.5

volts, VDC

0.5

-0.5

-1

0.5

1.5

2.5

3
-5

3.5

time, seconds x 10

Fig. A8.14 Plot created by Program Lab 0 RLC Discharge Data and
Model

wn = 6.293E5
z = 0.258
K = 3.29E-6
%
% Define "s" to be a transfer function in order to use
% s as the Laplace variable.
%
s = tf('s')
%
% Assign the transfer function to the variable "model".
%
model = K*wn^2/(s^2+2*z*wn*s+wn^2);
%
% Calculate the unit impulse of the model at the
% times in the vector time.
%
modelResponse = impulse(model,time);
%
% Determine the number of elements in the observed
% response data.
%
N = length(time)
%
% Calculate the time step dt.
%
dt = time(2) - time(1)
%
% Preallocate vectors for the variables to be
% 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.
%
error = zeros(1, N);
intError = zeros(1, N);
intAbsError = zeros(1, N);
intSqError = zeros(1, N);

%
% MATLABs lowest vector or array index is 1, not 0.
% Backward- looking recursive
% calculations cannot start with the index 1, since
% they would look back to the non-existent index
% value of 0. Calculate the first value of the error
% and integrals of the error, absolute value of the
% error, and of the error squared before the for end
%loop.
%
error(1) = modelResponse(1)-data(1);
intError(1)= error(1)*dt;
intAbsError(1)= abs(error(1))*dt;
intSqError(1)= error(1)*error(1)*dt;
%
% The for end loop uses backward-looking
% recursion to calculate the remaining values.
%
for i = 2 N;
error(i) = modelResponse(i) - data(i);
intError(i) = intError(i-1) + error(i)*dt;
intAbsError(i) = intAbsError(i-1) + abs(error(i))*dt;
intSqError(i) = intSqError(i-1) + error(i)*error(i)*dt;
end;
%
% A variables name followed by the Enter key is the
% MATLAB command to evaluate and display the
% result. We can display a result on the screen
% without format by entering the variable or elements
% name. To make our results more readable, we
% use the fprintf( ) function write to a file, but we omit
% the File_Variable_Name aka fid. When the fid is
% omitted, MATLAB interprets the function fprintf( )
% as a command to write to the screen.
%
fprintf('\nThe integral of the error
= %12.6E\n',intError(N))
%
% The syntax to continue a command onto the next
% line is three dots
%
fprintf('The integral of the absolute value of the ...
error = %12.6E\n',intAbsError(N))
%
fprintf('The integral of the error squared ...
= %12.6E\n',intSqError(N))

References and Suggested Reading


Hamming RW (1973) Numerical Methods for Scientists and engineers,
2ndedn. Dover, New York
Press WH, Teukolsky SA, Vetterling WT, Flannery BP (2007) Numerical recipes, 3rdedn. Cambridge University, Cambridge

Transfer Functions, Block Diagrams,


and the s-Plane

Abstract

Transfer functions are inputoutput relationships in the Laplace-domain. They are multiplicative operators. Multiplying a transfer function by the Laplace transform of an input variable yields the corresponding output variable. A block diagram represents inputoutput relationships 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


An operator is a function or an operation, in which one or
more input variables yield a single output variable. Operators are inputoutput relationships. We will restrict our discussion to single inputsingle output operators. Hence, we
will not consider multivalued functions to be operators. By
our definition, the quadratic equation is not an operator,
because it yields two values. Neither is multiplication of
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
test, that doubled input yields doubled output. Trigonometric
functions are not linear operators, since they are non-linear
functions.

9.1.1 Linear Operators


The simplest example of a linear operator is the operation
of multiplying a variable by a constant, say, the constant K.
We are all familiar with linear inputoutput relationships
in the time-domain, such as the relationship between the

displacement and force of a linear spring, where K is the


spring rate or spring constant
F (t ) = Kx(t )
(9.1)
Alternatively, we can express K as a multiplicative operator, by creating the ratio of the output variable, F(t) over the
input variable x(t)


Output F (t )
=
= K Operator
Input
x(t )

(9.2)

The next step is to generalize the familiar linear equation,


Eq.9.1, and express it in a functional form, Eq.9.3. Although it seems awkward, we will find this form useful for
extending our definition of operators from the time-domain
to the Laplace-domain


Output F (t )
=
= K{
Input
x(t )

} Operator

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_9, Springer Science+Business Media New York 2014

(9.3)

519

9 Transfer Functions, Block Diagrams, and the s-Plane

520

We need three more linear operators to demonstrate the


properties of linear operators. Let us add the value of 1,000
and the units, N/m, to the operator, K

K{

} 1, 000

Example: where u1 (t ) = [ ft ] and u 2 (t ) = [in ]


N
m
m

u 1 (t ) + 0.00254
u 2 (t )
0.305

m
ft
in
N
m
N
m
= 1, 000 0.305
u 1 (t ) + 1, 000
0.00254
u 2 (t )
m
ft
m
in

1, 000

N
m

Commutative property: L {K {u (t )}} = K {L {u (t )}}

Define the operators, L , F , and I as


Output Flb (t )
lb
=
= 0.224 = L {
Input
F (t )
N

} Operator

Example: where [u (t ) ] = [ m ] :
0.224

L {1 N} = 0.224 lb
and
Output
x(t )
m
=
= 0.305 = F {
Input
x ft (t )
ft

} Operator

F {1 ft} = 0.305 m
Output
x(t )
m
=
= 0.00254 = I {
Input
xin (t )
in

These linear operators possess one more important property, that of inversion, where the inverse of an operator is
indicated by the exponent, 1. For example, the inverse of
1
linear operator F { } is F { } . The property of inversion is
defined as


} Operator

Linear operators have associative, distributive, and commutative properties. To review these properties
Associative property: A ( B C ) = ( A B) C
Distributive property: A ( B + C ) = A B + A C
Commutative property: A B = B A

G{

{ {

}} = L {(F {K {u(t )}})}

N
1, 000
m

N
u (t ) = u (t )
m

Output
=3
Input

Let H be the operation of differentiating the input signal with


respect to time

Example: where [u (t ) ] = [ m ] :

H{

lb
m
N

0.305 1, 000
u (t )

N
ft
m

d
dt

Then

lb
m
N

= 0.224 0.305 1, 000 u (t )

N
ft
m
Distributive property:

H {G {u (t )}} =

d
(3u (t )) = y (t )
dt

and
G { H {u (t )}} = 3

}} = K {F {u (t )}} + K {I {u (t )}}

K F u1 (t ) + I u 2 (t )

1, 000

The multiplicative, linear operators, F { } , I { }, K { }, and


L { }, can be manipulated algebraically, as if they were constant factors, since, in fact, each is the operation of multiplying an input by a constant. We will state that all linear operators can be manipulated, as if they were constant factors,
because they all possess the four properties illustrated above.
We will not prove this statement, but we will illustrate it. To
illustrate this, let us define two linear operators. Let G be the
operation of multiplying input signal u(t) by 3

Associative property:

L F ( K {u (t )})

(9.4)

Example of the property of inversion

K 1 {K {u (t )}} =

9.1.2 Properties of Linear Operators

{ { } {

F { F {u (t )}} = u (t )
1

I {1 in} = 0.00254 m

0.224

N
N
lb
lb
1, 000 u (t ) = 1, 000 0.224 u (t ).
N
m
m
N

du (t )
= y (t )
dt

9.1 Linear Operators and Transfer Functions

521

These are equivalent statements since

b0

d
du (t )
(3u (t )) = 3
dt
dt

9.1.3 Incrementally Linear Functions


The operation of multiplying the input by a constant is clearly a linear operator. Although straight lines are described by
linear functions, only straight lines which pass through the
origin pass our test for linear operators. For example, the familiar equation of a straight line

d 2u
2 + b1
dt

du

L dt + b L {u}
2

d 2y
+ a1
2

dy

L dt + a L { y}

L dt

= a0

b0 s 2U ( s ) + b1 sU ( s ) + b2U ( s ) = a 0 s 2Y ( s ) + a1 sY ( s ) + a 2Y ( s ) .
2. Factor out the transformed input and output variables

(b s
0

+ b1 s + b2 U ( s ) = a 0 s 2 + a1 s + a 2 Y ( s )

m x(t ) + b = y (t )

3. Create a ratio of the output variable over the input variable

is not a linear operator, since doubling x(t) does not double


y(t)

2
Output ( s ) Y ( s ) b0 s + b1 s + b2
=
=
Input ( s ) U ( s ) a 0 s 2 + a1 s + a 2

m 2 x(t ) + b 2 y (t )
This equation of a straight line also does not yield an operator which is the ratio of output y(t) over input x(t)
Output y (t ) mu (t ) + b
b
=
=
= m+
Operator
Input
x(t )
x(t )
x(t )
A straight line which does not pass through the origin is referred to as incrementally linear, Sect.4.2.4.1.

9.1.4Differential Equations and Transfer


Functions as Linear Operators
The differential system equation is an inputoutput relationship. It is a dynamic operator which operates on the input
to yield the output, where output is the response function of
the dynamic system. A differential system equation is also
a linear operator, if each of its terms passes the double the
input yields double the output test. Linear differential system equations result from modeling the energetic properties
of a real object or system, using linear elemental equations.
Recall that we can create a transfer function from a differential system equation, as follows:
1. Use the Laplace transform, neglecting the initial condition terms, to create an algebraic equation
b0

d 2y
dy
d 2u
du
b
u
a
+
b
+
=
+ a1
+ a2 y
1
2
0
2
2
dt
dt
dt
dt

d 2u

du
+ b 2u = L
b 0 2 + b1
dt
dt

d 2y

dy
+ a 2 y
a 0 2 + a 1
dt
dt

(9.5)

A transfer function is a linear operator in the Laplacedomain. The numerator polynomial contains the operations performed on the input, transformed into the Laplacedomain, where the Laplace variable, s, assumes the role of
differentiation, with respect to time in the time-domain. Correspondingly, the denominator polynomial, a 0 s 2 + a1 s + a 2,
contains the operations performed on the output variable in
the differential equation. Note it is also the characteristic
function of the differential equation. Setting the characteristic function equal to zero forms the characteristic equation of
the differential equation

a 0 s 2 + a1 s + a 2 = 0
(9.6)
A transfer function is an unusual function, because it is used
in two very different ways. A transfer function can be evaluated as a conventional function by assigning a value to its
argument, s = + jw . Transfer functions are complex functions, meaning, in the general case, the argument, s, and the
output are complex numbers.
The second manner of using a transfer function is as a
dynamic operator. Rather than substituting the input function into the time-domain differential equation, the Laplacedomain transfer function acts as an inputoutput relationship, when multiplied by the Laplace transformation of the
input function. A transfer function is a multiplicative operator in the Laplace-domain. Multiplying a transfer function by
the Laplace transform of input function yields the Laplace
transform of the output (response) function.
Refer to the relationship between the displacement and
force of a spring, F (t ) = Kx(t ) . If we rearrange this as a ratio
of output F(t) over input x(t), we have the form of a transfer
function
F (t ) = Kx (t )

F (t )
=K
x (t )

9 Transfer Functions, Block Diagrams, and the s-Plane

522

A transfer function is a linear, multiplicative operator. For


example, if the input is x(t ) = 3 sin(2t ), then the output is the
product
x (t )

F (t )
x (t )

= K 3sin ( 2t )

F (t ) = K 3sin ( 2t )

If we perform the Laplace transformation on the time-domain


equation, where the unknown functions in the time-domain,
x(t) and F(t), are transformed into the unknown functions,
X(s) and F(s), in the Laplace-domain, and then create the
ratio of the output function over the input function, we have
a Laplace-domain transfer function:

L {F (t )} = L {Kx (t )}
F (s)

X (s)

F ( s ) = KX ( s )

=K

The advantage of a transfer function is its ability to serve as a


multiplicative constant, such as spring constant K. It can also
serve as a dynamic operator, meaning it can be formed from
the Laplace transformation of a differential system equation.
The time dependencies represented in the differential system
equation, as derivatives with respect to time, are represented
in the transfer function as powers of the Laplace variable s.

9.2Laplace-Domain Solution of a Set of State


and Output Equations
The final topic in our development of matrix mathematics is
use of the Laplace transformation to solve a set of state equations and output equations expressed in matrix form. The resulting solution is a vector of transfer functions of each of the
output variables. The solution eliminates the state variables,
unless they are included in the output vector. We begin with
state equations in standard vector-matrix notation

where C is the output matrix, and D is sometimes called


the feed-through matrix. The output equations are derived
from the equation list, presuming knowledge of the values of
the state variables over the duration of interest, and assuming
knowledge of the input, which is necessary for any solution.
We will solve the two matrix equations
dx
= Ax + Bu
dt
y = Cx + Du
by first solving the state equations, so that state vector x is
expressed in terms of the constant matrices, A and B, and the
input vector, u. We will then substitute the solution for x into
the output equations, yielding output vector y as a function
of input vector u.
There are only two steps in this solution, which require
mathematical insight or, for most of us, education. The first is
the fundamental approach to the solution. We will transform
the set of simultaneous first-order differential equations into
a set of simultaneous algebraic equations, by operating on
both sides of the state equations with the Laplace transform
dx

L dt = L {Ax + Bu}
L {y} = L {Cx + Du}
The Laplace transform is a linear operator and, as such, can
be manipulated as if it were a multiplicative constant. It can
be distributed to each term in a vector or matrix, in the same
manner that it is distributed to each term in a summation
dx

L dt = L {Ax} + L {Bu}
L {y} = L {Cx} + L {Du}

dx
= Ax + Bu
dt

Matrices A, B, C, and D can be factored out of the Laplace


transforms, since they are constants

where x is the vector of state variables; A is the coefficient


matrix comprised of terms consisting of the elemental parameters; B is the input matrix, again comprised of terms
consisting of the elemental parameters; and u is either a scalar, 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 derivation of the solution, then evaluate the solution, using a single
(scalar) input for u.
The output equations in matrix form are

L dt = AL {x} + BL {u}

y = Cx + Du

dx

L {y} = CL {x} + DL {u}


Recall the Laplace transformation of a derivative with respect to time yields the Laplace variable, s, in the Laplacedomain. Mathematical convention is to use capital letters for
the transformed variables
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

Considering the transformed state equations, subtracting


AX(s) from both sides yields
sX( s ) AX( s ) = BU( s )
This is where the second mathematical insight, or education,
is required. We would like to factor transformed state variable X(s) out of the two terms on the left side of the equation,
but we cannot, because the result

( s A ) X( s) = BU( s)

Y( s ) = CX( s ) + DU( s )
We can now substitute for the transformed state vector X(s)
in the output equation
Y( s ) = C ( sI A ) BU( s ) + DU( s )
1

Factoring out transformed input vector U(s), we obtain the


desired result

(s A)
which is the difference between the scalar variable s and the
matrix A. That operation is not defined. This term cannot be
evaluated. The mathematical insight is that we may pre- and
post-multiply any vector or matrix we wish by the identity
matrix, I, without changing the equation. If we pre-multiply
the transformed state vector, X(s), in the first term on the
left side
sIX( s ) AX( s ) = BU( s )
We have changed the equation no more than if we were to
have multiplied a term in a scalar equation by the scalar identity, the number one. However, we are now able to factor out
the transformed state vector, X(s)

( sI A ) X( s) = BU( s)
(9.7)
( sI A )

( sI A )1 ( sI A ) X( s) = ( sI A )1 BU( s)

v1g = LC

d 2 v3 g
dt

dx
= Ax + Bu
dt

dt

+ v3 g

d
dt

iL L
=
v
3g 1
C

1
i
1
L L
+

L v
v3 g

0
0

0
iC 1
0
i 1

0
0
R
iL
i = 1
0 + 0 v

v3 g

0 0
v12 R
v23 R 1
1

Recall the inverse of a matrix M is


cof ( M )

X( s ) = ( sI A ) BU( s )
1

Returning to the set of state equations and output equations

1 0 L
sI A = s

0 1 1
C

dv3 g

and output equations, Sect. 7.6.4

yielding

+ RC

and to a set of state equations, Sect. 7.6.3

M 1 =

X( s ) = ( sI A ) BU( s )

(9.8)

the transformed output vector, Y(s), as a function of the


transformed input vector, U(s).
The crux of the calculation is the inverse matrix ( sI A) 1 .
We will investigate the meaning of this expression by returning to the RLC circuit, Fig. 1.39, which we reduced in
Sect. 1.5.2.2 to the second-order system equation relating the
input, v1g, to the output, v3g,

y = Cx + Du

is the difference between two square matrices of the same


size. We solve for X(s) by multiplying both sides by the inverse of the factor, (sIA),

Y ( s ) = C ( sI A ) B + D U ( s )

contains the factor

since the factor

523

where cof (M ) is the matrix of the cofactors of M and |M| is


the determinant of M.
Expressing ( sI A) 1 in terms of the determinant and cofactor of ( sI A)

1
R

s 0 L
L

=
0 s 1
0

R
1
s +
L
L
=
1
0
C

1
L

9 Transfer Functions, Block Diagrams, and the s-Plane

524

( sI A )

R 1

s + L L
cof

1
T
s
sI A ]
C
[

=
=
R 1
sI A
s+
L L
1
s

coefficients of the exponents, which form the homogeneous


solution

(9.10)
v3 g H (t ) = A1e s1t + A2 e s2t
The homogeneous solution represents the natural response
of the system to a disturbance which transfers energy to the
system.
The transfer function, Eq.9.10, formed from the Laplace
transform of the differential system, Eq.9.8, contains the operators from the output variable side of the differential equation as the denominator polynomial

We will begin by evaluating the determinant


R
L
sI A =
1

1
L

s+

R
R
1

1 1
sI A = s + s = s 2 + s +

L C
LC
L
L
This expression is a second-order polynomial in s. It is, in
fact, the characteristic function of the system, which we can
verify by comparing to the characteristic function of the
RLC circuit we derive from the second-order system equation. Taking the Laplace transform of both side and neglecting the initial condition terms,
v1g = LC

d 2 v3 g
dt

L {v } = L LC
1g

L {v } = L
1g

+ RC

d 2 v3 g
dt

d 2 v3 g
LC
+
dt 2

dv3 g
dt

+ RC

L RC

+ v3 g

+ v3 g
dt

dv3 g

dv3 g
+
dt

L {v }
3g

V1g ( s ) = LC s 2V3 g ( s ) + RC sV3 g ( s ) + V3 g ( s )


V1g ( s ) = ( LC s 2 + RC s + 1)V3 g ( s )
R
1
1

V1g ( s ) = s 2 + s +
(9.9)
V3 g ( s )

L
LC
LC
What does it mean to have the characteristic function of the
system in the denominator of ( sI A) 1? Recall from the
Method of Undetermined Coefficients, Sect.2.5, when the
characteristic function of a system is set equal to zero, it is
the systems characteristic equation. The roots of the characteristic equation are the eigenvalues of the system. They are

Output ( s ) V3 g ( s )
=
=
Input ( s ) V1g ( s )

1
LC
R
1
s2 + s +
L
LC

(9.11)

Hence, the denominator of a transfer function is the characteristic function of the system.
Although the state-space representation of a dynamic system
is quite different from the same system represented as a higherorder differential equation, both mathematical representations
must contain the same physical information. From a mathematical perspective, invariant quantities do not change, when
we manipulate our mathematical representations. In dynamic
systems, the eigenvalues, or roots of the characteristic equation, and the corresponding eigenvectors, the set of orthogonal
vectors which correspond to the eigenvalues, do not change,
when we manipulate the mathematical representation, because
they have physical meaning. They are, respectively, the decay
rate and initial magnitude of the natural response of a system.
Turning to the calculation of the cofactor of ( sI A) , recall 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 substituting it in place of the element, Sect.7.5. Calculation of
the cofactor of a two by two matrix requires use of the special case of the determinant of a single element, since there
is only one element remaining in the matrix, when we drop
the row and column containing our target. Recall the determinant of a single element is defined as that element. Also,
recall the elements of the cofactor of a matrix are multiplied
by ( 1)( r + c ) = ( 1)(i + j ) . Hence,
R

s + L
cof ( sI A ) = cof
1
C

(1+1)
s
( 1)

cof ( sI A ) =

( 2 +1) 1
( 1)

L

1
L

1

C
R
( 1)(2+ 2) s +
L

( 1)(1+ 2)

9.2 Laplace-Domain Solution of a Set of State and Output Equations

Transposing the cofactors of ( sI A) yields

cof ( sI A )

s
=
1
L

s
C
=

R
1
s+

C
L

after we have taken the Laplace transform of output vector


y(t) for the RLC circuit,
1
L
.
R
s+
L

L {y } = L

Division by the determinant sI A yields the inverse matrix

( sI A )1

525

R 1

s + L L
cof

1
T
s

[ sI A ] = C
=
R 1
sI A
s+
L L
1

s
C

L {u} = L {v}

1
1


s
s
L
C

1 s + R
1 s + R

L
L = C
= L
R
R
1
1
2
2
s + s+
s + s+
L
LC
L
LC

Y ( s ) = C ( sI A ) B + D U ( s )
1

U (s) = V (s)

L
0
IC (s) 1

I (s) 1
1
R
0
+
s
R

C
L
I ( s ) = 1
0
1

2 R
0 s + s+
V12 ( s ) R
L
LC
V23 ( s ) R 1

0
0
1
L + 0 V ( s )

0 0
1

It will be easier to evaluate the right side, if we factor the


characteristic function, which is a scalar function, out of its
matrix product. We will divide the result of the matrix multiplication by the characteristic function as the last step. Factoring the scalar function to the front of its term

We can complete the solution by substitution




I C (s)
I (s)
R
Y( s) = I ( s)

V12 ( s )
V23 ( s )

and its scalar input, u(t),

( sI A )1

iC

iR
i
v
12
v23

(9.12)

0
IC (s)
1
I (s)
1
0 s
R

1
I (s) =
1
0

s2 + R s + 1
1
0
R
V12 ( s )
L
LC
C
V23 ( s )
R 1

0

1
1 0
L
L + 0 V (s)
R
s + 0 0

We now evaluate the matrix multiplication

IC ( s)
s

I ( s)

R
1

s
I ( s) =

s2 + R s + 1
V12 ( s )
L
LC
Rs
V23 ( s )

1
Rs

L

1 0
1
L + 0 V ( s )


L
0 0
R

1
L

R
R
s+

L
L

9 Transfer Functions, Block Diagrams, and the s-Plane

526

0
IC (s)
s


I ( s)
L

0
R
1
1

I ( s) =
+ 0 V ( s )
s

L

s2 + R s + 1
0
V12 ( s)
L
LC
R

1
s
V23 ( s)

1
s

LC
L

2 R

Ls + s +

L
LC

L s2 + R s + 1

L
LC
I
s
(
)
0
C

I (s)
s
0
R

I ( s ) = L s 2 + R s + 1 + 0 V ( s )
L
LC


0
V12 ( s )

Rs
V23 ( s )
1
R
1

Ls + s +

L
LC

R
1

L
LC

1
2 R

s + L s + LC


R
1

L s2 + s +

L
LC

1
2 R
L s + L s + LC
IC (s)

I ( s)
s

I ( s ) = L s 2 + R s + 1 V ( s )
L
LC


V12 ( s )

Rs
V23 ( s )

1
2 R
L s + L s + LC

R
1

L s LC

+ 1

s2 + R s + 1


L
LC

The final result is a vector of transfer functions for the output


variables of the system.
When there is the need to derive transfer functions of
higher-order systems, the vector-matrix form of the result,Y( s ) = (C( sI A) 1 B + D)U( s ), Eq.9.12, is a tremendous time-saving method. As a rule of thumb, the time required to reduce the energetic equations of a system to a
single, higher-order system equation is roughly proportional
to the square of the number of equations. In contrast, the
time required to reduce the same set of energetic equations
to state equations is roughly proportional to the number of
equations.
Modern handheld engineering calculators, i.e., the TI-89
and TI-Nspire, can invert a four by four matrix with symbolic elements. The symbolic algebra capability of these machines is superior to that of Mathcad and MATLAB. What
formerly were onerous linear or scalar algebraic expansions
and simplifications can now be performed quickly and accurately with these calculator.

9.3 Block Diagrams


Thus far, we have developed two mathematical descriptions
of the dynamic relationships present in energetic systems: (1)
classical, higher-order differential equations, which represent the dynamic relationship between a single input variable
and a single output variable, and (2) the state-space method,
using a set of simultaneous first-order differential equations,
expressed in vector-matrix notation and written in terms of


R
1

L s2 + s +

L
LC

IC (s)
s

V ( s)

L s2 + R s + 1

I R ( s )
L
LC

V ( s)
I ( s) 2 R
1

= Ls + s +

L
LC
V ( s)

V12 ( s )
Rs


V ( s) L s 2 + R s + 1

V ( s )
L
LC

23

V ( s ) R s 1

L
LC
+ 1

s2 + R s + 1


L
LC

9.3 Block Diagrams


Fig. 9.1a Block with a linear
operator. The input and output
signals conform to the direction
of the arrowheads. b Block with
transfer function G(s), input U(s)
and output Y(s)

527

b
Input

Linear Output
Operator

the energy storage, or state variables, and an input vector.


Linear operators permit both higher-order differential equations and sets of simultaneous differential equations to be
represented graphically as block diagrams. Block diagram
representation of differential equations is advantageous, because it yields a deeper understanding of the symbolic mathematics. Block diagrams have the additional importance as a
means of designing feedback control systems.
A block diagram is a method of representing information
flow, and the sequence of operations performed on information. The term, block, refers to the rectangles used to boxin the operators. Block diagrams are used for some engineering software graphical user interfaces. The drag-and-drop
diagrams you may have created using the data acquisition
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
signals, because the block diagram notation was developed
by electrical engineers. The lines representing signals must
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
have a single input and a single output. Do not show more
than one signal entering a block, except for a summation
junction. Never show more than one signal exiting a block.

9.3.1 A Block
The simplest block diagram contains a single block with
one line entering the block, the input signal, labeled U(s),
and one line exiting the block, the output signal, labeled Y.
The operator in the block is identified by the letter, G(s),
but does not need to be a function of the Laplace variable,
s. G(s) can be a constant. Since transfer functions are linear
operators, they can be used in block diagrams. Depending
on circumstances, we may choose to name the transfer function, Eq.9.13, G(s); express it as Y(s)/U(s), or as a ratio of
polynomials in s.
b0 s 2 + b1 s + b2
Y (s)
(9.13)
G (s) =
U (s)
a 0 s 2 + a1 s + a 2
The functional relationship represented by the block diagram
Fig.9.1b is expressed symbolically as

U(s)

G(s)

Y(s)

G {U ( s )} = Y ( s )
The output signal of the block can be labeled as a variable
(signal), Y(s), as the result of operation on the input, G {U ( s )},
or, with both terms, as an equation, G {U ( s )} = Y ( s ). The
only use of an equal sign on a block diagram is to indicate a
signal has two names.
Because linear operators can be manipulated algebraically, as if they were constant factors, we will simplify and
clarify our notation by omitting the braces, {}, which we
used in Sect.9.1.1. We may indicate a linear operator is operating on a quantity, by using the same notation as one does
for multiplication. For example, the distributive property expressed in Sect.9.1.1 as

{ { } { }} = K {F {u (t )}} + K {I {u (t )}}

K F u 1 (t ) + I u 2 (t )

will henceforth be expressed as

K FU 1 ( s ) + I U 2 ( s ) = KFU 1 ( s ) + KI U 2 ( s )
We may also omit the independent variable, s, and rely on
upper case font to identify variables in the Laplace-domain.

K ( FU1 + I U 2 ) = KFU1 + KI U 2
For convenience, we will apply the associative, distributive,
and commutative properties to combinations of signals and
operators. In other words, the variable (signal) need not be
on the right end of a term

LKI U = LKU I = LU KI = U LKI


When we interpret expressions, we must distinguish between
variables (signals) and operators. Operators operate on variables (signals), not the other way around.

9.3.2 Cascaded Blocks


Two operators in series are referred to as cascaded. Using
U(s) as the input signal, W(s) as an intermediate signal, Y(s)
as the output signal, and G(s) and H(s) as the two operators,
the cascade is shown in Fig.9.2.

9 Transfer Functions, Block Diagrams, and the s-Plane

Fig. 9.2 Cascaded blocks G(s)


and H(s)

U(s)

G(s)

Signal

W(s)
Signal

Linear
Operator
Fig. 9.3 Four equivalent block
diagrams

Fig. 9.4a Integration block with


Laplace-domain notation in the
time-domain. b Integration block
with time-domain notation

dv3g

dt

The functional relationship of the cascade is represented


symbolically as


UG = W and WH = Y

(9.14)

The commutative, associative, and distributive algebraic


properties of linear operators extend to their representation
in block diagrams. The order of cascaded blocks can be interchanged, or the block diagram simplified, by combining
the cascaded blocks. Block diagrams are equivalent, if an
input signal yields the same output signal. The four block
diagrams shown in Fig.9.3 are equivalent. The two cascaded
blocks simplify to the product of the two linear operators,
G(s) and H(s), as given in Eq.9.15:


UGH = Y

(9.15)

The term cascade derives from a natural cascade of water.


If one were to affect the water at a point in a cascade, say,
by disturbing the bottom and muddying the water, the effect
of that action would be seen downstream but not upstream.
The same assumption is made, when blocks are cascaded
in a block diagram, that actions downstream do not affect the
values of signals upstream. Whether or not this is a reasonable model in a dynamic system depends on the input and
output impedance of the components or subsystems which
comprise the blocks. Impedance is a generalized resistance,
addressed in Chap.11. A block with high input impedance
draws little power from its upstream neighbor. A block with

Signal

Linear
Operator

b
1
__
s

Y(s)

H(s)

v3g

dv3g

dt

GH

GH

528

dt

v3g

low output impedance can provide a large amount of power


to its downstream neighbor. Often, there is an impedance
mismatch in a system design, and the cascade model cannot
be used. In these cases, the dynamic model between the input
and the output variables must be represented as a whole by a
single transfer function.

9.3.3Differentiation, Integration, and Transfer


Functions Blocks
The example above illustrated how differentiation with respect to time is a linear operation. Integration and differentiation are inverse operations, since
du

dt dt = u

and

d
dt

( udt ) = u

There are different notations used to represent the operations


of differentiation and integration in time-domain (where the
signals are variables of time) block diagrams. Rather than
d
( ) and ( ) dt , some authors use D to represent differendt
tiation, and D1 to represent the inverse operation of integra1
tion. We will use s to represent differentiation, and s1 or
s
to represent integration in the time-domain, Fig.9.4. They
are concise and their use in the time-domain will reinforce
their meaning in Laplace-domain block diagrams.

9.3 Block Diagrams

529

9.3.4 Summation Junctions


Summation junctions are represented as circles, rather than
rectangles. A summation junction usually has multiple inputs, but a single input is allowable. There is only one output. A positive or negative sign is shown where the input
enters the summation junction. This allows the summation
junction to add orsubtract an input from the resulting output. Engineering terminology here is somewhat awkward.
The signal is described as inverted or non-inverted.
The advantage of this nomenclature is the reduction of potential sign errors, when a negative signal is inverted. Input
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
negative one as its operator), after the output signal leaves
the summation junction. Cascaded summation junctions are
interchangeable.

9.3.5 Branch Points


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
to provide additional copies. The signal does not divide at
a branch point, because what is transmitted is information.
The same information is sent down to both branches. The
graphical convention is to indicate a branch point with a dot,
in order to distinguish it from signal lines which simply cross
each other without a connection, Figs.9.5 and 9.6. These
conventions are identical to those used in circuit diagrams.
There is an important physical difference between branching
in a block diagram, and in an electric circuit. The quantity
which flows in a block diagram is information; whereas, the
flow quantities in an electric circuit are current and power.
In the abstract block diagram, there is no limit to the number
of branches, which can be added to an ideal signal. An ideal
signal never degrades, as it is divided.
A real signal in an electric circuit has a limit, called the
fan-out, which is the number of branches that can be added,
before the information of the signal degrades due to insufficient power. In a digital circuit, a typical fan-out is eight.
Fig. 9.6a One signal with
abranch point. b Two crossing
signals

Input B

Signal B

+
+

Output
A+B-C

Input C

A+B

A+B-C

A-C

+
+

A-C+B

Fig. 9.5 Equivalent block diagrams. a Single summation junction with


three inputs. bTwo summation junctions. c Summation junctions of b,
interchanged

9.3.6 Block Diagram Algebra


A set of mathematical rules is an algebra. Block diagram
algebra are the rules which permit manipulation of a block
diagram into an equivalent block diagram. We have already
introduced three rules:
1. Cascaded blocks can be interchanged or combined into a
single block.
2. A summation junction with three or more inputs can be
divided into cascaded summation junctions.
3. Signals can be branched to replicate signals.
We need two more rules to have a working knowledge of
block diagram algebra: how to move blocks and branch
points and how to collapse two different types of loops.
4. A block can be moved, relative to a branch point, if an
equivalent block diagram is created by adding or removing blocks, as necessary.
Rule four is best explained through two examples.

Signal B

Signal A

Input A

Signal A

b
Dot ties signals
together
Signal A

Signal A

Signal A

Signal A

9 Transfer Functions, Block Diagrams, and the s-Plane

530

G
A

Fig. 9.8 Two equivalent block


diagrams. a Block diagram with
block G after the branch point,
b Block diagram with a block
G before branch point and the
inverse block, 1/G, on the branch

Fig. 9.9 Feedforward loops

AG

G
H

AG+AH

A
A

G
H

Example 1 Move two identical blocks back (against the


direction of information flow) through a branch point. The
blocks merge (one is removed), Fig.9.7.
Example 2 Move a branch point forward through the block
G. When a branch point is moved forward through a block
G, the inverse block, 1/G, must be added to the branch to
remove the operation of block G on the signal, Fig.9.8.

9.3.7 Feedforward and Feedback Loops


We will state block diagram Rule Five and then develop the
background needed to understand it and apply it.
5. Loops must be untangled before they are collapsed from,
the inner most outward, using the feedback and feedforward loop reduction formulae.
Branch points and summation junctions allow block diagrams to include loops, where a signal is branched, operated on, and then summed. Forward and back in a

AG

AH

Information Flow

Fig. 9.10 Two equivalent block


diagrams. a A feedforward loop.
b The feedforward loop replaced
by a single transfer function

AG

Information Flow

AG+AH

++

AG
AH

1
__
G
A

Information Flow

AG

AG

AG

AG

++

Fig. 9.7 Two equivalent block


diagrams. a Block diagram with
two blocks G after branch point,
b Block diagram with a single
block G before branch point

Information Flow

AG
AH

++

AG+AH

G+H

AG+AH

block diagram are relative to the input and output signals.


The forward direction is toward the output of the block diagram. The backward direction is toward the input to the
block diagram.
Unfortunately, in the case of feedforward loops, liberty
is taken with the meaning of forward. Two feedforward
loops are shown in Fig.9.9. In Fig.9.9a, both signals of the
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
loop are in the same direction.
They can both be replaced by a single block with the same
transfer function, Fig.9.10. Replacing a loop with a single
transfer function is known as collapsing the loop.
A feedforward loop can be collapsed, i.e., replaced with
a single block, by inspection of the output of the summation junction in Fig.9.10a. Operating on input signal A with
the transfer function (the sum of the transfer functions of the
two branches of the feedforward loop, G+H) yields the same
output. Hence, the two block diagrams are equivalent.

9.3 Block Diagrams

531

G
H

Fig. 9.11 Negative feedback loop. Input R stands for the reference, and
output C stands for the controlled variable

Input R

Error E

Feedback CH

Output C

Fig. 9.12 Signals identified in a negative feedback loop

A feedback loop is shown in Fig. 9.11. This block


diagram is the basis of single inputsingle output feedback
control systems. The standard notation uses R for the input
signal, and C for the output signal, where R stands for Reference and C for the Controlled variable. The feedforward path
is from the summation junction, through transfer function G,
to the output. The feedback path is from the output, through
transfer function H to the summation junction. The feedforward transfer function, G, represents the system under automatic control, and includes the controller. The feedback
transfer function, H, represents the sensor used to measure
the output variable, C, and whatever signal processing is performed on it, before it is subtracted from the input signal at
the summation junction. Note the feedback signal is inverted
at the summation junction, creating negative feedback. The
output of the summation junction is the difference between
input R and feedback CH.The sensed output variable is error
signal E, Fig.9.12.
Suppose you wish to control the speed of a motor, using
closed-loop negative feedback control. Controlled variable
C would be the angular velocity of the motor. Input reference signal R could be a number of physical quantities, such
as the position of a throttle, the torque or force on a machine
element, a current, or a voltage. If reference signal R were
a voltage, then block H represents a tachometer which out-

Fig. 9.13 Signal names in block


diagrams. a An internal signal
with no name. b An existing
name, in this case, the output
signal of system C. c The name
error assigned to the output of the
summation junction

a
A

puts voltage. Why? The inputs to a summation junction must


have the same units, or they cannot be summed. The output of the summation junction, RCH, would be a voltage
which represents the error or difference between the desired
speed and the actual speed.
The transfer function, which is equivalent to a negative
feedback loop, cannot be determined by inspection of the
block diagram. The key to understanding block diagrams is
to identify the input and output signals of each block in the
diagram. There are three types of equations which can be
written to represent the inputoutput relationship of a block.
In many cases, the output of a block is an internal signal
within the system, which, though essential for the operation
of the closed-loop system, is not significant enough to be
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
system C, Fig.9.13b. This yields the non-trivial equation,
AF=C. The third case is an internal signal significant enough
to be named. Figure9.13c shows the summation junction of
a closed-loop negative feedback system. The output of the
summation junction is named error E. This yields the nontrivial equation, RCH=E.
To derive the single transfer function which can replace
or collapsed in a feedback loop, we need to identify the signals in the block diagram, write symbolic equations which
represent the inputoutput relationships of the blocks, and
do a little algebra. We will use Fig.9.12 where the output of
the summation junction is named E for Error. There are two
blocks and the summation junction, so we can write three
equations. The equation for the summation is
R CH = E
We can also write an equation for the controlled variable, C,
in terms of E
EG = C
The remaining equation is trivial, since the output signal is
named as the product of the input signal and the block, but
we will write it, to be comprehensive,

b
F

AF

c
F

Output
C=AF

Input
R

Error
E=R-CH

CH
Feedback

9 Transfer Functions, Block Diagrams, and the s-Plane

532
Fig. 9.14 Two Equivalent block
diagrams. a Block diagram of a
closed-loop negative feedback
system. b Equivalent transfer
function of the closed-loop
system

Input
R

Error
E

Control
Signal
U

G Plant

Output
C

Input
R

HSensor

CH = CH
These three equations can be reduced to yield an inputoutput relationship for the input and output signals, R and C, of
the system by eliminating the signal, E:
EG = C E =

C
G

R CH = E R CH =

G
_____
1+GH

GController

C
G
=
R 1 + GH

Control
Signal
U

G Plant

Output
C

HSensor

Open loop here


Fig. 9.16 Open-loop system. The feedback loop is broken before the
summation junction. The error signal, E, equals input R

increases, the feedback signal, CH, grows, and error E decreases.


An absolute requirement for closed-loop feedback control
shown in Fig.9.14 and Eq.9.16 is that a positive input to the
closed-loop system, R, produces a positive output, C. A sign
inversion in the forward path between the summation junction and the output leads to positive feedback.
R ( CH ) = R + CH E

C
G

RG CGH = C RG = C + CGH RG = C (1 + GH )
GC .L.

Error
E

Feedback
CH

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

G
H

GController

Feedback
CH

(9.16)

Equation 9.16 is known as the closed-loop transfer function since it replaces the entire feedback loop.
The block diagram of Fig.9.14a is the simplest and most
condensed block diagram of a closed-loop negative feedback
system. Transfer function G represents all subsystems in the
forward path combined. A slightly expanded block diagram
is shown in Fig.9.15. Feedforward transfer function G has
been split into two transfer functions, GController and GPlant,
where plant is control jargon for the system under control.
The controller uses error signal E to produce control signal U, which drives the plant. Suppose the plant was a DC
motor running in steady-state, when the input, or reference
signal R, was increased by a step change. The difference between input R and feedback signal CH, also known as error
E, is greatest at the instant the step command is given to the
system, but before the system has responded. As motor speed

Positive feedback is good for people but destabilizing for


physical systems. Positive feedback causes the error signal
to grow when, in fact, the system is acting to reduce the error.
The system is out of control.
Recall the sign inversion introduced by the linear graph
transformer and transducer sign convention. It was stated
that the signs would be fixed after the reduction of the
linear graph to a system equation, if necessary. Use of a
transfer function derived by the linear graph method in a
closed-loop feedback control system is such a situation.
There must not be a net sign inversion in the path from the
summation junction back to the summation, the open-loop
signal shown in Fig.9.16. If there is a sign inversion, then
it must be fixed either mathematically by simply removing
the negative sign or physically by inserting an inverter
into the signal path.
Figure9.16 shows the open-loop configuration, where
the feedback loop is opened, or broken, before the summation junction. If feedback signal CH is zero, then input R is
the only information available to the controller. Regardless
of the load on the motor, the controller sends the same control signal. Open-loop control is the least expensive, in terms
of hardware. However, the control algorithm is designed for

9.3 Block Diagrams

Input
R

533

Error
E

Control
Signal
U

G Plant

select a controller type, then determine its gains to optimize


the response of the system.

Output
C

HSensor

Feedback
CH

Fig. 9.17 Proportional closed-loop negative feedback control

the expected operating conditions, which are not necessarily


the actual conditions the system experiences.
The simplest closed-loop negative feedback controller is
a proportional controller, Fig. 9.17. A proportional controller multiplies error signal E by a constant, the controller
gain, K, to produce control signal U. Hence, the control signal is proportional to the error signal.
When a block diagram with a controller is reduced to a
single transfer function, the controllers gain, K, is kept as a
factor in the feedforward transfer function, Fig.9.18.
Recall the denominator of a transfer function is the transfer functions characteristic function, which, when set equal
to zero, yields the transfer functions characteristic equation.
The closed-loop transfer function:


GC .L. =

KG
C
=
R 1 + KGH

9.3.7.1 Collapsing Nested Loops


Feedback and feedforward loops are often nested within
one another. The simplest case is when none of the branches
of the inner and outer loops overlap. For example, the block
diagram in Fig.9.19a has two nested feedback loops.
Reduction of nested loops always proceeds from the inside out. This strategy applies to both feedback loops and
feedforward loops. Collapsing the inner loop using the reG
yields the block diagram
duction formula GC .L. ( s ) =
1 + GH
Fig.9.20.
The reduction formula is used again to collapse the outer
loop, Fig.9.21. The final step is to clear fractions. Clearing
fractions often requires more effort than the block diagram
algebra since the transfer function F, G, and H are ratios of
polynomials.
The final form of a transfer functions is a proper ratio,
Fig. 9.22. This result is in symbolic form. After the transfer functions F, G, and H are substituted, fractions must be
cleared again. The denominator is then expanded into a polynomial and factored.
Using the following transfer functions as an example,
F (s) =

(9.17)

4
s+6

1 + KGH = 0

(9.18)
C
=
R

The eigenvalues of the closed-loop system and, hence, the


closed-loop systems transient response are a function of the
controller gain, K. The process of control system design is to
Fig. 9.18 a Block diagram of
closed-loop negative feedback
control with controller gain K as
a factor. b Closed-loop systems
transfer function

3
s+9

b
KG

5
s+7

3
s+9
3 5 4 3
1+
+
s + 9 s + 7 s + 6 s + 9

a
R

H (s) =

C
G
=
R 1 + GH + FG

yields the characteristic equation of the closed-loop system




G (s) =

KG
______
1+KGH

H
Fig. 9.19 Nested feedback loops

Outer Loop
Inner Loop

9 Transfer Functions, Block Diagrams, and the s-Plane

534

G
_____
1+GH

The numerator is in factored form and can be left alone.


Expand the denominator

3 ( s + 6) ( s + 7 )
C
= 3
2
R
s + 22 s + 159 s + 378 + (15s + 90) + (12 s + 84)

Fig. 9.20 Inner loop collapsed

Collect the denominator into a polynomial

G
______
1+GH
_________
FG
______
1+

3 ( s + 6) ( s + 7 )
C
=
R s 3 + 22 s 2 + 186 s + 552

Factor the denominator into first-order and underdamped


second-order factors

1+GH

3 ( s + 6) ( s + 7 )
C
=
R ( s + 6.42) s 2 + 15.58s + 85.98

Fig. 9.21 Outer loop collapsed

G
_________

1+GH+FG

The TI-Nspire calculator function propFrac ( ) will evaluate


this transfer function symbolically and return a proper ratio.
The manual technique follows.
Multiply by the unity ratio

9.3.7.2 Untangling Loops


If loops overlap and are, therefore, entangled, then block
algebra must be used to untangle the loops before the loops
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
invested untangling entangled loops so that the loop reduction
formulae can be applied from the inner most loop outward.
The process of untangling loops requires a combination
of moving branch points, adding blocks, and dividing summation junctions. Moving summation junctions through a
block can be done, but it is more difficult than the other actions. Consider the block diagram in Figs.9.23 and 9.24.
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
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
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
1
Point B, then the inverse operation,
must be added to
H2
the Branch Point As branch to remove the effect of H2. Conversely, if the Branch Point B is slid backward through Block

Fig. 9.22 Transfer function equivalent to the block diagram in Fig.9.19

C ( s + 6)( s + 7 )( s + 9)
=
R ( s + 6)( s + 7 )( s + 9)

( s + 6)( s + 7)( s + 9)
( s + 6)( s + 7)( s + 9)

3
s+9
3 5 4 3
+
1+
s + 9 s + 7 s + 6 s + 9

Distribute and cancel common factors

( s + 6) ( s + 7)(3)
C
=
R ( s + 6) ( s + 7 ) ( s + 9) + ( s + 6)(3)(5) + ( s + 7 )( 4)(3)

Fig. 9.23 Block diagram with


entangled feedforward and feedback loops

H1
R

H2

G1
G2

++

9.4 Time-Domain Block Diagrams of Differential System Equations

535

Fig. 9.24 Block diagram with


entangled feedforward and feedback loops

H1

Branch
Point A

H2

G1

Branch
Point B
+
+

G2
Fig. 9.25 Branch Point A slid
forward through block H2 to the
right of Branch Point B

1
___
H2

Branch
Point B

H2

G1

H1
+
+

Branch
Point A

G2
Fig. 9.26 Branch Point B slid
backward through Block H2 to
the left of Branch Point A

Branch
Point B

G2

H2 to the left of Branch Point A, then the operation H2 must


be added to Branch Point Bs branch.
The loops can be collapsed once they are untangled. Remember to identify whether a loop is a feedforward or a feedback loop by the relative direction of the information flow.
A feedforward loop has the information flowing in the same
direction (possibly backward) on both sides of the loop parallel to the input-to-output direction. A feedback loop has information flow in both directions. Do not rely on the signs at the
summation junction to identify either type of loop.

9.4Time-Domain Block Diagrams


of Differential System Equations
A differential equation can be represented graphically as a
block diagram. First, the differential equation is manipulated
to represent the highest-order derivative of the output variable, as a summation of the other output variable terms plus
the input variable terms. Then, the highest derivative of the
output variable is integrated to yield the next highest derivative. The process is repeated by a series of integrators, until
the output of the last integrator is the output variable itself.

H2

G1

H1

H2

+
+

Branch
Point A

9.4.1Block Diagram Without Differentiation


of the Input
We will illustrate the technique with a third-order differential
equation, with u as the input variable, and y as the output
variable
b 3u = a 0

d 3y
d 2y
dy
a
+
+ a2
+ a3 y
1
3
2
dt
dt
dt

The first step is to put the differential equation into standard


form, by clearing the highest-order derivative of the output
variable of its coefficient
b3
d 3 y a1 d 2 y a2 dy a3
u
=
+
+
+
y
a
dt 3 a0 dt 2 a0 dt a0
0
where the constant terms are shown in parentheses. What are
the unknowns in this equation? The constants are known,
since they are terms comprised of element parameters. Input
u(t) is also known. We must know the input, i.e., what we
are doing to the system, if we wish to predict the response of

9 Transfer Functions, Block Diagrams, and the s-Plane

536
Fig. 9.27 A chain of integrators
yielding progressively lowerorder derivatives

d3 y
___
dt 3

d2 y
___
dt 2

1
__
s

Fig. 9.28 Integrators and coefficients of output variable terms

d3 y
___
dt 3
a3 ___
d2 y
__
a0 dt 2

Fig. 9.29 Time-domain


block diagram of
d 3y
d2y
dy
b3u = a 0 3 + a1 2 + a 2
+ a3 y
dt
dt
dt

b3
__
a0

+_ _ _

d3 y
___
dt 3

1
__
s
d2 y
___
dt 2

1
__
s
a3
__
a0

dy
___
dt

a2 ___
dy
__
a0 dt

1
__
s

dy
___
dt

1
__
s

a2
__
a0

d2 y
___
dt 2

1
__
s

a1
__
y
a0

1
__
s

dy
___
dt

1
__
s

a1
__
a0

1
__
s

a3
__
a0
a2
__
a0
a1
__
a0

the system. Consequently, only output variable y(t) and its


derivatives are unknown.
Now, express the differential equation as a summation,
yielding the highest-order derivative of the output variable
a1 d 2 y a 2 dy a 3
d 3 y b3
u
=

a dt 2 a dt a y
dt 3 a0
0
0
0
We integrate the highest derivative, d 3 y /dt 3 , with respect
to time to yield the second highest derivative, d 2 y /dt 2 . The
process is repeated by arranging the integrators in series,
so that the output of an integrator is the input to the next,
until the final integrators input is dy /dt and its output is
y (t ). The chain of integrators, Fig.9.27, produces the output
variable and all of its derivatives, allowing us to create the
summation.
The rule is simple: Never differentiate the output signal
or its derivative. Only integrate them. The output signals of
the integrators are then multiplied by a constant term and fed
back to the summation junction, Figs.9.28 and 9.29.

The time-domain block diagram representation of the differential equation can be used to define a set of state variables. Recall the standard form of state equations is a set of
first-order differential equations
dx
= Ax + Bu
dt
Remember that following the flow of a block diagram backward through a linear operator is equivalent to performing
the inverse operation represented in the block
State variables are defined on the block diagram, starting with the output variable. Define the output variable as
x1. The input to the final integrator is the derivative of x1 ,
dx1 /dt . This signal is given two names by also defining it
as the second state variable, x 2. Assigning two names to the
same signal creates an equation


dx1
dt

= x2

(9.19)

9.4 Time-Domain Block Diagrams of Differential System Equations

dy
__
dt

1
__
s

dy
__
dt

Fig. 9.31 Define state variables


starting with x1 y, where y is
the output variable, then work
backward through the integrators.
Define as many state variables
as there are integrators, i.e., the
number of state variables equals
the order of the differential
equation

dx3
dt

1
0
a2

a0

(9.20)

dt

dx2
dt

d2 y
dt 2

1
__
s

x2 =

dx1
dt

dy
dt

1
__
s

x1
y

State variables generally have no physical meaning. If the


output variable, x1, were displacement, its third derivative would be the derivative of acceleration, also known as
jerk. If the output variable were pressure, the only physical
meaning of its third derivative would be as the derivative
of the second derivative of pressure. State variables are not
unique. Since we use the linear graph method to extract state
equations from energetic systems, our state variables are the
energy storage variables of the system. However, once we
perform any matrix algebra on the state equations, say to
manipulate them into a form preferred for Modern Control
theory, we create a new set of state variables to describe the
same system.

9.4.2Block Diagram with Differentiation


of the Input
The third-order system equation we expressed as a block
diagram was a special case. There was no differentiation of
the input variable. The general case of a third-order system
equation is

x1 0

x +
u
a 1 2 b 3
x3
a 0
a0
0

(9.21)


x3 =

1
__
s

d3 y
dt 3

dx1
= x2
dt
dx2
= x3
dt
a3
b3
a2
a1
dx3
= x1 x2 x3 + u
dt
a0
a0
a0
a0

0
x1
d 0
x2 =
dt a 3
x3 a
0

dt

Define state variables working


BACKWARD through the integrators.

and, in matrix form




dy
__
dt

Remember there are only as many state variables as there


are integrators. Do not define the derivative of the final state
variable as a state variable, or you will define one too many.
The derivative of the final state variable is the output of the
summation junction, Figs.9.30 and 9.31.
The state equations are those equations, which can be
written for the inputs to each integrator. All but the final state
equation will define a state variable as the derivative of the
next one. The final state equation will be a summation, including all of the state variables and the input.
The state equations are

dy
__
dt

Fig. 9.30 Moving in the reverse


direction through a block inverts
the operation of the block.
Forward through an integrator, or
reverse through a differentiator,
is the same operation, because
the operations are inverses

537

b0

d 2u
d 2u
du
+ b1 2 + b 2
+ b3u
2
dt
dt
dt
d2y
d2y
dy
= a0 2 + a 1 2 + a 2
+ a3 y
dt
dt
dt

where the input variable can be differentiated up to the degree


of the output variable, but not higher. A system equation, in

9 Transfer Functions, Block Diagrams, and the s-Plane

538
Fig. 9.32 Block diagram of the
summation, Eq.9.23

b2
___
a0
du
__
dt
d2 u
___
dt 2

b1
___
a0
b0
___
a0

++
+ __

d2 y
___
dt 2
a1
___
a0

dy
__
dt
a2
___
a0

which the input variable has higher-order differentiation than


the output variable, describes a system which responds to
what the input is going to do, not what the input just did.
Energetic systems (without intelligence) do not anticipate
the future. The mathematical term for a system with higher
differentiation of the input than the output is non-causal,
meaning that it violates cause and effect.
Differentiation of the input function presents difficulties
for the state-space representation of a dynamic system, because state equations can only have derivatives of the state
variables, not of the inputs. There are two special cases
which can be implemented.
Case One: The input functions are known. In this case,
the derivatives of the input function can be evaluated analytically, and treated as additional inputs. This case is identical
to a system with multiple undifferentiated inputs.
Case Two: The input functions are known to be continuous and smooth. If the input function is continuous and
smoothly varying, then a finite difference approximation
of a derivative will be reasonably accurate. However, the
finite difference approximation of the derivatives of discontinuous yields meaningless results which are inversely proportional to the time step t.

9.4.2.1 Numerical Differentiation of the Input


We will illustrate the method using a generic second-order
differential equation, with u as the input variable, and y as
the output variable


b0

d 2u
dt 2

+ b1

du
d2y
dy
+ b 2 u = a0 2 + a 1
+ a 2 y (9.22)
dt
dt
dt

which we rearrange as a summation, equaling the highestorder derivative of the output variable

a1 dy a 2
d 2 y b0 d 2 u b1 du b2
= 2 +
+ u

y
2
 dt
a0 dt
a0 dt a0
a0 dt a0
(9.23)

The block diagram of the differential equation is constructed
by starting with the summation junction. The output of the
summation junction is the highest-order derivative of the
output variable, Fig.9.32.
The block diagram is completed, by adding differentiators to create the derivatives of the input variable, integrators
to create the output variable, and its derivatives, Fig.9.33.
Note there are two integrator blocks in the block diagram.
The number of integrators equals the order of the differential equation. Note the input signal terms are fed forward,
and the output variable terms are fed back. Although we
can implement this block diagram, if input u(t) is a continuous, smoothly varying function which can be differentiated
numerically, that is not the preferred form. The difficulty
lies in numerical differentiation being sensitive to the size
of time step t. If the input function changes smoothly, the
approximation of its derivative can be accurate. However,
when the function changes rapidly, the size of the time step
determines the value of the derivative.

9.4.2.2Variable Redefinition to Eliminate


Differentiation of the Input
It is possible to manipulate the differential equation to eliminate differentiation of input. Recall we introduced a method
of representing system equations with derivatives of the
input terms in Chap.7. The method was presented as an
algorithm symbolically (or formulaically) for expressing a
transfer function in Controllable Canonical Form. Recall
that canonical means standard. Controllable canonical

a1
___
a0

du
__
dt

d2 u
___
dt 2

b2
___
a0

b1
___
a0

b0
___
a0

++
+ __

d2 y
___
dt 2

1
__
s

dy
__
dt

a2
___
a0

1
__
s

9.4 Time-Domain Block Diagrams of Differential System Equations

Fig. 9.33 Block diagram of the second-order differential equation,


Eq.9.21, with differentiation of the input function

d4y

dt

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
controller gains to be applied to each state variable.
We will now derive the similar algorithm, to express a
higher order, differential system equation in Observable
Canonical state-space form. We will develop the method
graphically using a block diagram, and then extract the recursive algorithm from the block diagram. We will use the
same general fourth-order system equation used in Chap.7.

a0

d4y
dt

+ a1

d3y
dt

+ a2
= b0

d2y
2

dt
d 4u
dt

+ a3
+ b1

dy
+ a4 y
dt

d 3u
dt

+ b2

d 2u
dt

+ b3

du
+ b4 u
dt

The first step is to express the differential equation in standard form, by clearing the highest-order derivative of the
output variable of its coefficient.
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
b d 4 u b d 3u b d 2 u b du b4
u
+
+ 0 4 + 1 3 + 2 2 + 3
a0 dt a0
a0 dt
a0 dt
a0 dt
Now, isolate the highest derivative of the output variable on
the left-hand side of the equation (or, in other words, solve
for it)
a1 d 3 y a 2 d 2 y a 3 dy a 4
d4y
=

a dt 3 a dt 2 a dt a y
dt 4
0
0
0
0
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

Next, integrate both sides the number of times necessary, to


integrate the highest derivative output variable down to the
output variable, four times in this case

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
u dt dt dt dt
= 3 2
y+ 4 + 3 + 2 +
+
a0 dt a0
a0 dt
a0 dt
a0 dt a0
a0 dt
a0 dt
a0 dt

9 Transfer Functions, Block Diagrams, and the s-Plane

540

On the right-hand side, collect the input and output terms of


the same order of differentiation, and distribute the integrals

This integral equation represented as a block diagram in


Fig.9.34.

a d 3 y b d 3u
b d 4u
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
a4
a dy b3 du
b4
+ 3
+
dtdtdtdt + y + u dt dt dt dt
a0
a0
a0 dt a0 dt

Evaluate the integrations, neglecting the constants of integrations. (The integration constants represent initial condition terms, which will be satisfied when the set of state equations are solved.) All integrations will be performed on the
fourth-order derivative of the input and output variable. Only
three will be performed on the third-order derivative, leaving
one left over. Two will be performed on the second derivatives, leaving two left over, etc.
a1
b1
b
y = 0 u + y + u dt
a0
a0
a0
a2
b2
+ y + u dt
a0
a0
a3
b3
+ y + u dt dt dt
a0
a0
a
b
+ 4 y + 4 u dt dt dt dt
a
a0
0
The four integrals are terms of the input variable, and the
output variable integrated to increasing degrees. One way
to represent the flow of this calculation is to nest the
integrals. Integration proceeds from the innermost integral,
which, when evaluated, is summed with the next term. That
sum is then integrated, and the process proceeds

We can represent this fourth-order system equation


a0

d4y
d3y
d2y
dy
+
a
+
a
+ a3
+ a4 y
2
1
2
3
4
dt
dt
dt
dt
d 3u
d 4u
d 2u
du
= b0 4 + b1 3 + b2 2 + b3
+ b4 u
dt
dt
dt
dt

in state-space by extracting the state and output equations


from the block diagram. The first step is to identify the state
variables as outputs of the integrators, and derivatives of the
state variables as input to the integrators, starting at the right
with the first state variable, x1, and working toward the left.
Now, write equations for the output of each of the summation junctions, again, working from left to right
b
y = x1 + 0 u
a0
a1
b1
dx1
= x2 y + u
dt
a0
a0
a2
b2
dx2
= x3 y + u
dt
a0
a0
a3
b3
dx3
= x4 y + u
dt
a0
a0
a
b
dx4
= 4 y + 4 u
dt
a0
a0

b
y= 0u
a0

a
b1 a 2
b2
b3
b
a1
a 3
+ 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

Output y

b
y = x1 + 0 u
a0

+
+

a1
b b
dx1
= x2 x1 + 0 u + 1 u
dt
a0 a0
a0
a2
b b2
dx2
= x3 x1 + 0 u + u
dt
a0 a0
a0

x1

b0
a0

1
s

a3
b b3
dx3
= x4 x1 + 0 u + u
dt
a0 a0
a0

a1
a0

+
+

dx
1
dt

a
b b
dx4
= 4 x1 + 0 u + 4 u
dt
a0
a0 a0

Distribute and separate variables to yield the proper form of


the state and output equations:

x2

b1
a0

a2
a0

+
+

dx2

dt

1
s

b
y = x1 + 0 u
a0

1
s

b3 a 3 b
a3
dx3
= x1 + x4 + 0 u
dt
a
0
a0 a0 a0

x4

a3
a0

dx3

dt
+
+

b1 a1 b
a1
dx1
= x1 + x2 + 0 u
dt
a0
a0 a0 a0
b2 a 2 b
a2
dx2
= x1 + x3 + 0 u
dt
a
0
a0 a0 a0

x3

b2
a0
b3
a0

541

b a b
a
dx4
= 4 x1 + 4 4 0 u
dt
a0
a0 a0 a0

a4
a0

Input u

b4
a0

dx4

dt

1
s

The state and output equations expressed in matrix form are

Fig. 9.34 Block Diagram of Observable Canonical form

The first equation is the output equation, which, importantly,


allows us to eliminate output variable y from the state equations. Substitute for y in the state equations

a1

a0

x1 a 2


d x2 a0
=
dt x3 a 3

x4 a0
a
4
a0

b1 a 1 b0

1 0 0

a0 a0 a0

x b2 a 2 b
0 1 0 1 0
x2
a0 a0 a0
+
u
x3 b 3 a 3 b0
0 0 1
x
4 a0 a0 a0

b a b
0
4 4
0 0 0

a0 a0 a0

x1
x
b
2
y = [1 0 0 0] + 0 u.
x3 a0

x4

542

9 Transfer Functions, Block Diagrams, and the s-Plane

Reversing the order of the state variables, which interchanges the rows and columns of coefficient matrix A, puts the
state and output equations into Observable Canonical form:

x4
1

d x3
=
dt x2
0
x1

0 0
0 0
1 0
0 1

y = [0 0 0

b4 a4 b0
a
4

a0
a0 a0 a0

a 3 x b3 a b
4 3 0
a0 a0 a0
a0 x3
+
u
a 2 x2 b 2 a2 b0

a0 x1 a0 a0 a0

a1
1 a 1 b0

a0 a0 a0
a0

x4
x
b
3
1] + 0 u
x2 a0

x1

9.5State Equations as a Time-Domain Block


Diagram

v(t)

C
g

Fig. 7.6 Linear graph of the series RLC circuit of Fig.7.5

9.5.1Drawing a Block Diagram of an Existing


State Equation
If a set of state equations has been derived for a system, then
constructing a block diagram for the system using the state
equations is straightforward. We will use the energy storage
variables of this system, as its state variables. The state equations, derived in Sect.7.6.2.3 from using current through the
inductor iL as the energy storage variable and voltage across
the capacitor v3g, as the state variable, are
diL
1
1
R
= iL v3 g + v1g
dt
L
L
L
dv3 g 1
= iL
dt
C

Presented below are three methods to represent the state


equations of a system as a block diagram. One method is to
reduce the equation list of the state equations, then construct
the block diagram from the state equations. A second method is to draw the block diagram directly from the equation
list. This method can save time but requires experience. The
third method is to define a new set of state variables from a
higher-order system equation. The latter can also be used for
system equations, which do not have derivatives of the input
variable.
All three techniques will be illustrated, using the series
RLC circuit from Sects.1.4.2.2 and 7.6.2. The linear graph
and the energetic equations of the series RLC circuit from
Sect.7.6.2.1 are reproduced below, for reference.

Energetic Equations

9.5.2Drawing a Block Diagram


from the Energetic Equations

Compatibility: v(t ) = v12 + v23 + v3 g


Continuity: isource = iR
Elements: v12 = RiR

iR = iL
v23 = L

Energy: E system = E L + EC

diL
dt

iL = iC
iC = C

1
E L = LiL2
2

dv3 g
dt

1
2

EC = C v32g .

(7.45)

Note the state equation for iL is a summation, but the state


equation for v3g is not a summation.
The structure of a block diagram of state equations usually has as many summation junctions as there are state equations, since state equations are usually sums, Fig. 9.35. The
output signals of the summation junctions are the derivatives
of the state variables. The block diagram also has as many integrators as state equations, since the derivatives of the state
variables must be integrated to yield the state variables.

The method of drawing the state equations directly from the


set of energetic equations of system is not very practical, but
it is useful as an aid in understanding the parallels between
the symbolic mathematics and block diagrams. The method
follows the procedure for deriving the state equations from
the equation list. Start with the elemental equations for the
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


state equations, Eq.7.45

v
Input

1
__
L

di L

dt

dv3g
___
dt

v23

Inductor:
v23 diL
di
v23 = L L
=
(9.24)
dt
L
dt

dt

iC dv3 g
=
C
dt

iL iC = 0

iC

(9.25)

The block diagrams are built backward from the integrators.


We need to provide the signals fed into the integrators, in
terms of the input, v1g, and the state variables, iL and v3g.
Providing signal iC, in terms of the input and the state variables, 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
iC = iL

Providing signal v23 in terms of the input and the state variables requires a summation junction and a few blocks. Rearrange the compatibility equation to yield voltage v23
(9.26)
v = v12 + v23 + v3 g v v12 v3 g = v23
The input to the system is v, and v3g is a state variable. We
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
v12 = RiR
Current iR is not a state variable, but it equals the current
through the inductor iL, which is a state variable. Using the
continuity equation for node two with the element equation
for the resistor

1
__
s

iL

R
__
L

iL

1
__
C

iL

1
__
s

v3g

1
__
L

Capacitor:
dv3 g

v3g

of the state variables, will be the input to integrator blocks.


The integrators yield the state variables, Fig.9.36.

iC = C

1
__
L

1
__
C

iL
State Variable

v3g
State Variable

di L
___
dt
v23
___
L
dv
3g
___
dt
iC
__
C

1
__
s

iL

1
__
s

v3g

Fig. 9.36 Block diagrams yielding the state variables a the current
through inductor iL, from Eq.9.24 and b The voltage across the capacitor, v3g, from Eq.9.25

iR iL = 0

v12 = RiL

yields v12 in terms of state variable iL. Substitution of this result into Eq.9.26 creates an equation for voltage v23 in terms
of the input and the state variables


v v12 v3 g = v23

v RiL v3 g = v23 (9.27)

The block diagram of the state equations, Fig.9.38, is created by assembling the block diagram fragments of Figs.9.36
and 9.37.

9.5.3Drawing a Block Diagram of State


Equations from a System Equation
This method defines new state variables from the higherorder system equation. The block diagram of the state equations uses the new state variables. Our preference is to use
the energy storage variables of a system as its state variables.

9 Transfer Functions, Block Diagrams, and the s-Plane

544

v12
v12
v

We start the recursive definitions, using output variable v3g


as the first state variable, x1,

iL

iR

(9.28)
v3 g x1

v23

The recursion is to define the next state variable as the derivative of the previous state variable:

v3g

dv3 g

dt
Fig. 9.37 Block diagram of Eq.9.27

These state variables have physical meaning. They are


physical quantities that can be measured. The energy variables of a system do, indeed, define the energetic state of
the system. However, state variables are not unique. As we
saw in Chap.7, linear operations performed on a set of state
variables yield a new and valid set of state variables. Although alternative sets of state variables are mathematically
legitimate, they may have no direct physical meaning. Consequently, we will continue to emphasize that energy storage
variables are state variables. One important application of an
alternative set of state variables is to enable the expression of
a higher-order system equation in state-space.
Recall from Chap.7, that a higher-order system equation
without derivatives of the input variable can be expressed
in state-space, using a recursive definition of the state variables, based on the output variable. Let us again use the RLC
circuit example from Chaps.1 and 7. The system equation,
which relates input voltage v1g to voltage across capacitor
v3g, is this second-order differential equation


v1g v = LC

d 2 v3 g
dt

+ RC

dv3 g
dt

+ v3 g

(1.48)

Fig. 9.38 Block diagram of


the RLC circuit drawn from its
energetic equations

dx1
dt

x2

(9.29)

Substituting the state variables, x1 and x2, and their derivatives into Eq.1.49 yields two equations, either


v = LC

dx1
dx2
+ RC
+ x1
dt
dt

(9.30)

or
v = LC

dx 2
dt

+ RCx2 + x1

(9.31)

Which form do we use for a state equation? Equation 9.30


cannot be expressed as a state equation, because it has two
dx
dx2
derivative terms, 1 and
. A state equation can have
dt
dt
the derivative of only one state variable.
We will use Eq.9.31 where we have made the substitution
dx1
= x 2 ,to eliminate the extra state variable derivative. The
dt
substitution necessary to eliminate the extra state variable
is the first state equation, Eq.9.29. Rearranging Eq.9.31 to
express the derivative of state variable x2 as the sum of products of constant terms, state variables and input, yields the
state equation:
dx2
1
1
R
=
x1 x 2 +
v1g
dt
LC
L
LC

v12
v12
v

(9.32)

iL

iR

v23

1
__
L

iC

1
__
C

v3g

di L
___
dt
v
23
___
L

1
__
s

iL

dv
3g
___
dt
iC
__

1
__
s

v3g

9.6 The s-Plane

dx 2
___
dt

545

x2

1
__
s

x1

1
__
s

dx1
___
dt

Fig. 9.39 Block diagram of Eqs.9.33 and 9.34. Integrators yielding


the state variables, x1 and x2

The state equation for x 2 , Eq.9.32, is a summation of the


terms of input v and state variables, x1 and x 2 ,. Fig. 9.40. The
complete block diagram is formed by adding gain blocks to
multiply input v and state variable x1 by 1/LC, and state variable x 2 by R/L, Fig. 9.41.

9.6 The s-Plane


1
___
x
LC 1

+
-

dx 2
___
dt

R
___
x
L 2
Fig. 9.40 Block diagram of Eq.9.31, the state equation for x2

Reordering and summarizing the state equations, Eqs.9.29,


9.32, and the output equation, Eq.9.28,
dx1

dt
dx 2

= x2
=

dt
v3 g = x1

R
1
1
x1 x 2 +
v
LC
L
LC

(9.33)

The block diagram is constructed, starting with the output of


the summation junction, which is the derivative of the second state variable. We need a cascade of integrators in series
to yield the state variables, x1 and x2, Fig.9.39.


dx 2
dt

dt = x 2 =

dx1

(9.34)

dt

dx1
(9.35)
dt dt = x1

We will plot values of the roots of the characteristic equation, the eigenvalues of the system. Since the roots may be
complex, we will use a complex plane with the real axis,
, and the imaginary axis, j. 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 understood that observed frequency would never be the ideal,
undamped, never-seen-in-nature, natural frequency, n.
Although we commonly refer to the complex plane,
it is correct to refer to a complex plane. Every complex
function is an inputoutput relationship associated with two
complex planes, one for the input and one for the output.
Complex variable s of the s-plane is the variable, introduced
when we solved system equations, using the method of undetermined coefficients. The s-plane is the plane of the roots
of the characteristic equation of a system equation. These are
the characteristic values or eigenvalues of the system. The
symbol is generally used when eigenvalues are derived
from the coefficient matrix A of the state-space representation of a system. The symbol s is used when the eigenvalues
are calculated from the characteristic equation derived from
the differential system equation or the transfer function.
In the general case, s is a complex number, s = + jw .
The values of s are the exponents of the homogeneous solution. If s is a purely real number, s=, the exponentials are
real exponentials decays. If s is a complex number, then the
homogeneous solution is underdamped and may oscillate. In
that case, the homogeneous solution is the product of a real
exponential and the sum of two exponentials with imaginary
exponents, j, which represent the oscillation.

Fig. 9.41 Block diagram of Eq.1.48

1
x1
LC

1
x1
LC
v
Input

LC

1
v
LC

R
x2
L

dx2

dt

LC
dx1

dt
x2

s
R

x2

x1

x1
y

Output

9 Transfer Functions, Block Diagrams, and the s-Plane

546

We will use a second-order system with a damped, oscillatory homogeneous response as the example. The general
form of the characteristic equation is
s 2 + bs + c = 0 s 2 + 2 n s +  n2 = 0
Solving the characteristic equation

(2w n )2 4w n2

b b 2 4c 2w n
=
s1 , s2 =
2
s1 , s2 =

2w n 4 2w n2 4w n2
2

2w n 2 2w n2 w n2
2

s1 , s2 = w n 2w n2 w n2
The damping ratio must be less than one for the homogeneous response to be oscillatory. This leads to complex roots

< 1 2w n2 < w n2 2w n2 w n2 < 0.


Multiplying both terms by 1 in the form of j2 reverses the
order of the terms in the difference

2w n2 w n2 = j 2w n2 j 2 2w n2
and yields the familiar form when
of the radical

j 2 = j is factored out

s1 , s2 = w n 2w n2 w n2 = w n

j 2w n2 j 2 2w n2

s1 , s2 = w n j w n2 2w n2 = w n jw n 1 2
Matching real and imaginary terms
s1 , s2 = j w
s1 , s2 = w n jw n

= w n


1
j w = jw n 1 2
2

Note that the real component is negative, because the


damping ratio and the ideal, undamped natural frequency
n are both positive. The negative value of the real component results from our assumed of a damped, or decaying, homogeneous response. The imaginary term is not
the natural frequency. It is the damped natural frequency,
which is the frequency one observes. The name, natural frequency, is unfortunate, because the frequency it refers to is
completely undamped, which is not natural. A better term
would be the ideal or undamped frequency. However,
the term, natural frequency, is the nomenclature. For systems
with damping ratio of 0.1, the difference between the observed, damped frequency and the ideal, undamped, natural
frequency is 1% or less.

9.6.1Poles, Zeros, and Pole-Zero Transfer


Function Form
A useful form of a transfer function, and the form we use
to design control systems, is to express the numerator and
denominator of a transfer function as the product of factors,
or, in other words, in factored form. The factored form of the
denominator allows us to see the roots of the characteristic
equation, the systems eigenvalues, by inspection. The roots
of the numerator are important in the design of control systems, but a discussion of their effects requires introduction to
the root locus, a topic in classical control theory.
Equation9.36 is a second-order transfer function expressed in factored form


G (s) =

Output b0 s 2 + b1 s + b2 K ( s + z1 )( s + z2 )
= 2
=
( s + p1 )( s + p2 )
Input
s + a1 s + a2

(9.36)
The roots are the opposites of the constants, i.e.,
s = p1 and s = p2 are the roots of the denominator, and
s = z1 and s = z2 are the roots of the numerator.
Thus far, we have used transfer functions as multiplicative operators. It is also possible to evaluate them as conventional functions by substituting a complex number for
the variable, s = + jw . The result of evaluating a complex
function for an argument is a complex number, which can be
expressed in Cartesian form, polar form, or as the product of
its magnitude and a complex exponential unit vector.
G ( s ) = G ( + jw ) = x + jy = G ( + jw ) e jG ( + jw )
(The terminology of evaluating a transfer function is awkward, because the angle of a complex number is its argument, the term used for the value operated on by a function.)
The magnitude of the transfer function, G(s), evaluated
for s = z1 or s = z2 is zero, since the magnitude of the
numerator is zero.
G ( z1 ) = G ( z2 ) = 0
However, the magnitude of the transfer function, G(s), evaluated for s = p1 or s = p2 , is infinite, since the magnitude
of the denominator is zero.
G ( p1 ) = G ( p2 ) =
The values, s = z1 or s = z2 , are called zeros of the
transfer function, G(s), since they yield zero magnitude.
The values, s = p1 and s = p2 , are called poles of G(s).
The term, pole, comes from a three dimensional plot, with
the components of s = + jw forming the two axes of a
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


magnitude increases sharply, as the value of s approaches
s = p1 or s = p2 , from any direction in the s-plane. That
shape is typically described as a spike, but the Laplace
variables symbol is s, so a different term, pole, is used.
Example The transfer function, G(s), expressed as the ratio
of two polynomials and in pole-zero (factored) form
G (s) =

G ( s1 ) = G ( 2 + j 2)
G ( 2 + j 2) =

-8

j4

<0

j2

-6

-4

UNSTABLE

>0
2

-2

10

-j2
-j4
-j6

Fig. 9.42 The right-half of the s-plane represents unstable systems, because the eigenvalues have a positive real component , which leads to
unbounded growth of the homogeneous, or natural, response

nitude for values of s along the line = 2. Figure9.53b


shows the magnitude for values of s along the line j =0,
which is the real axis.

9.6.2Stability

352

( 2 + j 2 + 2) ( 2 + j 2 + 4) ( 2 + j 2 + 8)

G ( 2 + j 2) =

352
( j 2) ( 2 + j 2) ( 6 + j 2)
o

G ( 2 + j 2) = 8.80 j 4.40 = 9.84e j153


Tabulating the results
G ( s) =

-10

j6

Stable

352
352
Output
= 3
=
Input
s + 14 s 2 + 56 s + 64 ( s + 2)( s + 4)( s + 8)

Evaluate G(s) for s1=2+j2, s2=2+j, s3=2+j0.5,


s4=2+j0.25, s5=2+j0.1, and s6=2+j0.05. Present the
result in Cartesian and complex exponential form.
Evaluating a transfer function in the manner of a conventional function by substituting for the argument s, the
value of a complex number can be done on an engineering
calculator

s-plane

Output
352
352
= 3
=
2
Input
s + 14 s + 56 s + 64 ( s + 2)( s + 4)( s + 8)

G ( 2 + j 2) = 8.80 j 4.40 = 9.84e j153

G ( 2 + j ) = 15.2 j 20.9 = 25.9e j126

G ( 2 + j 0.5) = 18.3 j 53.7 = 56.7e j109

G ( 2 + j 0.25) = 19.2 j115 = 116e j 99.5

G ( 2 + j 0.1) = 19.5 j 292 = 293e j 93.8

G ( 2 + j 0.05) = 19.5 j 586 = 586e j 91.9

Note the increase in the magnitude of the result from 9.8 to 586,
as the value of s approaches the pole at s=2 in the negative
imaginary direction, from s1 = 2 + j 2 to s6 = 2 + j 0.05,
forms a spike in magnitude at the pole. It is easier to interpret these results graphically, rather than as a list of complex
numbers. Figure9.53a shows the transfer functions mag-

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
equilibrium by a change to the systems input. Specifically,
we are interested in how quickly energy is dissipated. If the
system oscillates due to internal energy flows, what is the
frequency of the oscillation?
The natural response is known as the unforced response, which is somewhat of a misnomer, because the natural response is produced, when the forcing function (also
known as driving function, or input) is turned on or off, or
when it changes form, such as from a step to a ramp. Both the
impulse and step inputs reveal the homogeneous response of
a system, and are used as test inputs. The impulse response is
the natural or homogeneous response, since the particular solution for an impulse response is zero. Although the impulse
inputimpulse response relationship has great theoretical
utility, since any input can be represented by the integration
of impulses, in practice, the step response is easier to work
with due to power limitations. The steady-state response to
the step input must be subtracted, if it is non-zero, to yield
the homogeneous response.
The engineering adage, all passive systems are stable,
means that all systems without feedback control are inherently stable. This is not completely true, since it is possible to
construct an unstable passive system, but they destroy themselves the first time they are operated. Thus the adage applies
to all systems which have been used once.
Unfortunately, it is not only possible but common for a
stable, passive system to become de-stabilized, when placed

9 Transfer Functions, Block Diagrams, and the s-Plane

548

A
____
s+a

Input

A
____
s+a

Faster Decay

s-plane
j2

Output

Fig. 9.43 Block diagram of a unit impulse input to a first-order system,


and the resulting unit impulse response of a first-order system

-12 -11 -10

-9

-8

-7

-6

-5

-4

-3

-2

-1

-j
-j2

under closed-loop feedback control. Consequently, the most


important design criterion for a control system is stability,
which is assessed by the position of the poles of the closedloop system on the s-plane.
Positive values of the real component produce unstable
responses. Hence, the roots of a closed-loop systems characteristic equation must be in the left-half of the s-plane for the
response to be stable. The imaginary axis is the boundary between the stable left side and the unstable right side, Fig. 9.42.
The reason that the stability of a system is determined by
the real component is clear from the solution of the system equation. Recall the homogeneous solution of a secondorder system is
yh (t ) = A1e 1 + A 2 e 2 = A1e( + j w )t + A 2 e( j w )t
s t

s t

yh (t ) = A1e t e j wt + A 2 e t e j wt = e t A1e j wt + A 2 e j wt

where the undetermined coefficients, A1 and A2, depend on


both the system and the initial conditions
y (0+ ) = I .C.1

and

dy (0+ )
= I .C.2
dt

Expressing the sum of the complex exponentials,


( A1e jt + A2 e jt ), as the trigonometric function
A1e jt + A2 e jt = A sin(t +  )
by using one of Eulers equations
cos ( ) =

e j + e j
2

or

sin ( ) =

e j e j
2j

Fig. 9.44 Purely real poles, s=. The distance from the imaginary axis
is the decay rate,

e a =

1
ea

Consequently, when the real component is negative, such


as = w n , the real exponential decays with increasing
time t
1
e n t = n t
e
whereas, if the exponent is real, then the real exponential
increases with increasing time.
Physically, a system becomes unstable, because energy accumulates in the system over time. A stable system dissipates
energy. If the homogeneous response of the system is oscillatory, the accumulation of energy in the system increases the
amplitude of the oscillations, as the energy flows between
independent, internal energy storage elements. Eventually,
if the power supplied to the system is sufficiently large, the
amount of energy stored in the system exceeds the capacity
or strength of the system, thereby destroying it. In reality,
many systems of interest to mechanical engineers have limited power supplies. These systems will reach a bounded unstable state, in which their unstable responses reach a finite
limiting value.
To use the s-plane to design controllers, we must be able
to associate the position of a closed-loop real pole or complex
conjugate pair of poles on the s-plane with the contribution
that pole (or those complex conjugate poles) make(s) to the
time-domain response of a system. A convenient method for
doing so is to plot the impulse responses. Recall the Laplace
transformation of a unit impulse is unity

f (t ) = (t ) = for t = 0

f (t ) = (t ) = 0 for t 0
F (s) = 1

Amplitude A of the sinusoid is constant. However, the purely


real exponential multiplying the sinusoid is time-varying.
Hence, the real exponential factor governs whether the amplitude of the response grows or decays.
Recall a negative exponent represents a fractional and a
sign inversion:

Consequently, in the Laplace-domain, the impulse response


and the transfer function have the same form, although they
are different physical quantities, since the response is a variable (or signal), and the transfer function is an inputoutput
relationship (or dynamic operator), Fig. 9.43. It is essential
that the history of a Laplace-domain calculation is known.
The resulting ratio of polynomials in s cannot be assigned
meaning, if the history is unknown.

yields

e t A1e j wt + A 2 e j wt = e t A sin (w t + )



Sinusoid with
amplitude A

549

A
_____________
2
s + 2ns + 2n
A
_____________
Output
s 2+ 2ns + 2n

1.0
0.8

-10t

0.6

-5t

-2t

-t

Input

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

t, sec

Fig. 9.45 Impulse responses of first-order systems with poles, s= 1,


s= 2, s=5, and s= 10

9.6.3 Real Component , the Decay Rate


Restricting our discussion to the left-half of the s-plane, the
real exponential controls the rate of decay of the homogenous response. A fast response decays quickly. Again, the
negative real exponential is interpreted as
e a =

1
ea

Therefore, the more negative (i.e., further left on the negative real axis) the real component of s = + jw , the faster is
decay of the response Figs. 9.44 and 9.45. The time-domain
response, corresponding to a purely real value of s, is nonoscillatory. The relationship between the value of and the
speed (or decay rate) of the homogeneous response is illustrated with their impulse responses. Note that all of the
responses are first-order decays with the same shape, but
scaled in the time dimension. If a system is stable, its homogeneous response must decay with time. Hence
e t = e

and

where time constant

The actual amplitude contributed by a single real pole in a


higher-order transfer function would depend on the coefficient determined by the partial fraction expansion, known
as the residue, which, in turn, is governed by the transfer
function evaluated for the magnitude of the pole, s = , after
the factor, ( s + ( )), was removed. This fact will be illustrated below, when we consider the effect of relative position
on the magnitude of contribution of each of three poles to the
response of a system.

9.6.4Imaginary Component , the Observed,


Damped Frequency
The imaginary component of the closed-loop pole,
s = + jw , is the observed frequency of the homogenous

s-plane

j11
j10

Increasing Frequency

Impulse Response

9.6 The s-Plane

j9
j8
j7
j6
j5
j4
j3
j2
j
-5

-4

-3

-2

-1

Fig. 9.47 Complex eigenvalues with decay rate, =2, and increasing
frequency . Only the complex pole in the upper half-plane is shown,
of the conjugate pairs of poles

response, which is the damped natural frequency, Fig. 9.47.


For convenience, we do not use the subscript, d, to indicate
that the of the imaginary component, j, is the damped
frequency, jd. It is understood, because all physically systems have damping due to either to mechanical friction, fluid
viscosity, or electrical resistance.
Varying the imaginary component, j, of a closed-loop
pole, while holding the real component, , constant produces homogeneous responses of varying shapes, as shown
below. The constant decay rate, , creates a fixed duration
for the various responses. Increasing the frequency, j, increases the number of oscillations which changes the shape
of the response. We illustrate this with impulse responses in
Figs. 9.48 and 9.49.
Note the peak amplitude of response increases with increasing frequency, since the sine function reaches its first
maximum earlier at higher frequency.

9.6.5 Damping Ratio


The ideal, undamped natural frequency derives from the
standard form of an oscillatory second-order transfer function factor

9 Transfer Functions, Block Diagrams, and the s-Plane

550

Impulse Response

0.8

s-plane

e-2t sin(10t)

0.6

e-2t sin(5t)

0.4

0.0

0.0

0.5

1.0

t, sec

1.5

2.0

Fig. 9.50 The s-plane, the complex plane used to plot the eigenvalues
of dynamic systems. One of a complex conjugate pair of eigenvalues
is shown

Fig. 9.48 Underdamped second-order impulse responses


Table 9.1Factor

Impulse Response

1.0

e -2t

0.5

1 2

reduces n to  d =  n 1 2

e-2t sin(5t)

0.0
-0.5
-1.0
0.0

e-2t sin(50t)
0.5

1.0

t, sec

1.5

2.0

Fig. 9.49 Underdamped impulse responses and their decay envelop


for systems with s = 2 + j 5 and s = 2 + j 50

G ( s) =

N (s)
s + 2w n s + w n2
2

This form is useful, because the damping ratio, , and period


of a damped oscillation, T, can be determined from an impulse or step response of a second-order system. The damping ratio is calculated using the log decrement formula.
The damped period yields the damped frequency, d
2 radians
=
T
second

w wd =

(9.37)

The damped frequency yields the natural frequency, n




-0.2
-0.4

jd = jn 1- 2

e-2t sin(2t)
e-2t sin(t)

0.2

s = + j

d = n 1 

n =

d
1 

(9.38)

What we call the natural frequency, n, is, in fact, not natural but an ideal, which would occur in the absence of energy
dissipation. The discrepancy between the actual, damped fre-

1 2

0.707

0.707

0.5

0.866

0.4

0.917

0.3

0.954

0.2

0.980

0.1

0.995

0.05

0.999

quency, or d, and the ideal, undamped, natural frequency


varies with the damping ratio, .
Table 9.1 illustrates the difference between the actual,
observed, damped frequency, or d, and the ideal, undamped natural frequency, n, is certainly negligible, when
the damping ratio 0.1.
Closed-loop poles which fall on a ray from the origin of
the s-plane have the same damping ratio, . This is established using Pythagoreans theorem, to determine the distance from the origin to the closed-loop pole, which is the
ideal, undamped natural frequency, 1, Fig.9.50.

(w )
n

+ wn 1 2

= 2w n2 + w n2 1 2

2w n2 + w n2 w n2 2 = w n
The inverse cosine of the angle , which the ray from the
origin makes with the negative real axis to find the damping
ratio, , is shown in the s-plane construction, Figs.9.50 and
9.51.
cos ( ) =

w n
=
wn

(9.39)


Homogeneous oscillatory responses with the same damping
ratio, , have similar shapes, Fig. 9.52, but these are compressed or extended along the time axis, because the damping ratio, , is a factor in the decay rate, = w n, and the
damped frequency, w d = w n 1 2 .

9.6 The s-Plane

551

0.6

Co

j11

Impulse Response

s-plane

an
nst

j10
j9

tD

j8

am

j7

pin

j6

gR

j5

atio

j2

-4

0.0

0.0

1.0

-3

-2

-1

2.0

t, sec

3.0

4.0

Fig. 9.52 Impulse responses of underdamped second-order systems


with the same damping ratio, . Notice the shape of the curves is the
same, except for the temporal scaling

j
-5

e-2t sin(4t)
e-t sin(t)

j3

-6

e-3t sin(6t)

0.2

-0.2

j4

e-5t sin(10t)

0.4

Fig. 9.51 Complex eigenvalues with constant damping ratio 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 = 
highest power of s of the denominator polynomial. The nuUsing Eq.
 = 9.39,
63.4. cos(63.4 ) = 0.45 = 

merator and denominator are then factored.

9.6.6s-Plane Plots of Transfer Function Poles


and Zeros
A plot of a transfer functions poles and zeros conveys most
of the information of the transfer function in graphical form.
Recall that poles and zeros derive their names from the magnitude of the transfer function evaluated as a conventional
function for a value of s. The magnitude spikes when the
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
constant, gain K, which is a factor of the numerator. The gain
cannot zero the numerator, except in the trivial case of setting K=0. Consequently, the gain is not a zero. The transfer
functions gain, K, is not represented in a plot of the poles
and zeros since it is not a function of s. The converse is true,
as well. A transfer function can be constructed in factored
form from its pole-zero plot, except for the unknown gain, K.
The s-plane provides a graphical design tool for feedback control system design. For the time being, we will use
it to plot the roots of the denominator and the numerator of
transfer functions, after they have been factored into polezero form. The conventional symbols are Xs for poles, roots
of the denominator, and Os for zeros, roots of the numerator.

9.6.7 Example Pole-Zero Plots


Poles and zeros are the roots of the denominator and numerator polynomials, respectively, of a transfer function. The first
step in plotting the poles and zeros of a transfer function in
polynomial form is to write the transfer function in standard

Example One: Express the transfer function, G1(s), in standard form and plot its poles and zeros (if any) on the s-plane,
Fig. 9.54.
G1 ( s ) =

704
352
= 3
2 s + 28s + 112 s + 128 s + 14 s 2 + 56 s + 64
3

G1 ( s ) =

352
( s + 2)( s + 4)( s + 8)

(9.40)

Example Two: Express the transfer function, G2(s), in


standard form and plot its poles and zeros (if any) on the
s-plane, Fig.9.55.
 G (s) =

816
272
= 3
2
3s + 48s + 252 s + 480 s + 16 s + 84 s + 160
3

G2 ( s ) =

272
( s + 4 j 2)( s + 4 + j 2)( s + 8)

(9.41)

Example Three: Express the transfer function, G3(s), in


standard form, and plot its poles and zeros (if any) on the
s-plane, Fig.9.56.



G3 ( s ) =

14 s 2 + 168s + 728 14
s 2 + 12 s + 52
=

0.5s 3 + 5s 2 + 9s + 8 0.5 s 3 + 10s 2 + 18s + 16


G3 ( s ) =

28( s + 6 j 4)( s + 6 + j 4)
( s + 1 j )( s + 1 + j )( s + 8)

(9.42)

9 Transfer Functions, Block Diagrams, and the s-Plane

552
Fig. 9.53 Magnitude
of the transfer function,
G ( s ) = 352 ( s + 2) ( s + 4)( s + 8),
as the value of s approaches the
pole at s=2. a Section along
line = 2. b Section along real
axis, =0

600

400

|G(s)|
200
0
-j2

-j1.5

-j

-j0.5

j0.5

j1.5

j2

Imaginary Component of s = -2 + j

3,000

2,000

|G(s)|
1,000

0
-3

-2.75

-2.5

-2.25

-2

-1.75

-1.5

Real Component of s = + j0

s-plane

s-plane

j3

j5

j2

j4

j3
j2

-10 -9 -8

-7

-6 -5 -4

-3 -2

-1

-j

-10 -9 -8

-j2

-7 -6

-5

-4 -3 -2

-1

-j

-j3

-j2

Fig. 9.54 Pole-zero plot of G1(s), Eq.9.40

-j3
-j4

s-plane

-j5
j4

Fig. 9.56 Pole-zero plot of G3(s), Eq.9.42

j3
j2
j

-10 -9 -8

-7 -6 -5

-4 -3 -2

-1

1
-j
-j2
-j3
-j4

Fig. 9.55 Pole-zero plot of G2(s), Eq.9.41

Summary

Linear Operators
The important properties of linear operators are that they can
be distributed and their output scales with their input. These
properties allow them to be used with superposition and as
operators in block diagrams. All blocks in a block diagram
are a single inputsingle output relationship, except summation junctions which can have any number of inputs but only
one output.

Problems

553

Time-Domain Block Diagrams


Integration and differentiation, inverse operations, are more
conveniently represented by their Laplace-domain symbols
than by their time-domain expressions. Higher-order differential equations can be expressed as time-domain block diagrams in which a chain of integrator blocks yields progressively lower-order derivatives of the output variable. Block
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 representing system equations with differentiation of the input in
state-space are more complicated due to the need to redefine
the state variables so as to eliminate the differentiation of
the input.

Transfer Functions and Laplace-Domain Block


Diagrams
Transfer functions created from differential system equations are used as operators in block diagrams. A block diagram with a single input and a single output can be reduced
to a single transfer function. The properties of transfer functions derived by reducing block diagrams are the same as
those derived directly from differential system equations.
Specifically, the denominator of the transfer function is
the characteristic function of the system represented by the
block diagram and, when set equal to zero, is the characteristic equation. Closed-loop feedback control is described by
a block diagram in which the output variable is fed back and
subtracted from the input to yield the error signal. The roots
of the denominator of a closed-loop system transfer function
are the systems eigenvalues.

s-plane
The roots of the numerator and denominator of a transfer
function are called the zeros and poles of the system, respectively. The magnitude of the transfer function is zero when
evaluated for a value of s which is a root of the numerator.
Roots of the denominator, i.e., the eigenvalues, yield an infinite value of the magnitude of the transfer function. Plots
of the eigenvalues on the s-plane show their position relative
to the real and imaginary axes and to one another. Since the
eigenvalues determine the homogeneous or natural response
of a system, the s-plane is used as a graphical design tool.

v(t)

Fig. P9.1 RLC circuit

Problems
Problem 9.1 An electric circuit consisting of a voltage
source, a resistor, R = 1 , an inductor, L=2H, and a capacitor, C=3F, is shown in the schematic, Fig.P9.1. The
circuit is de-energized for time, t<0.
9.1.a Formulate the state equations for this system and
express them in standard vector-matrix form.
9.1.b Determine the eigenvalues, the ideal, undamped natural frequency n, the damped natural frequency d,
and the damping ratio .
9.1.c Use the Laplace transformation to solve the state
equations for a vector of the transfer functions of the
state variables.
9.1.d Plot the poles and zeros (if any) of the transfer functions for the voltage across the capacitor and the current through the inductor on the s-plane.
9.1.e Assuming the circuit is initially de-energized, use
Laplace transforms and partial fraction expansion, if necessary, to determine the voltage across
the capacitor and the current through the inductor, when the input voltage applied to the circuit is
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.
Problem 9.2 A translational mechanical system consisting
of a mass M sliding on a lubricating fluid film with damping
b is shown in Fig.P9.2. A spring K is attached between the
mass and ground. The system is at rest and relaxed, before
the mass is acted upon by an applied force F(t). Two sets of
parameter values are tabulated in Table P9.2.
Table P9.2 Parameter values
M

Case I

Case II

9 Transfer Functions, Block Diagrams, and the s-Plane

554

x,v

F(t)

P(t)

R1

Pump

vent to
atmosphere

R2

Fluid
Reservoir

Lubricating fluid
Damping b
Fig. P9.2 Force source acting on a spring-mass-damper system

Shaft rigidly attached


Hydrodynamic
bearing damping b2
Flywheel
rotational inertia J
Compliant shaft
torsion spring K
Fluid coupling
damping b1
Angular velocity
input (t)

Fig. P9.3 A rotational mechanical system

9.2.a Derive the state equations of this system and express


them in vector-matrix form.
9.2.b Calculate the systems eigenvalues for Cases I and
II. If the system is underdamped, determine the ideal,
undamped natural frequency n, the damped natural
frequency d, and the damping ratio .
9.2.c Use the Laplace transform to solve the state equations
to yield the transfer functions for the state variables
X 1 (s)
X (s)
and 2 .
U (s)
U (s)
9.2.d Use Laplace transforms and partial fraction expansion,
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
MATLAB.
Problem 9.3 The rotational mechanical system shown in
Fig.P9.3 consists of an angular velocity (t) which drives
a compliant shaft with torsion spring constant K through a
fluid coupling with damping b1. The shaft drives a flywheel
with inertia J. The hydrodynamic bearing of the flywheel are
modeled as a drag cup with damping b2.
Table P9.3 Parameter values
J

b1

b2

Case I

20

50

0.2

Case II

200

0.5

Fig. P9.4 A fluid system

9.3.a Derive the state equations of this system.


9.3.b Calculate the systems eigenvalues for Cases I and II
using the parameter values in Table P9.3. If the system is underdamped, determine the ideal, undamped
natural frequency n, the damped natural frequency
d, and the damping ratio .
9.3.c Use the Laplace transform to solve the state equations
to yield the transfer functions for the state variables
X 1 (s)
X (s)
and 2 .
U (s)
U (s)
9.3.d Use Laplace transforms and partial fraction expansion, if necessary, to determine the responses of
the state variables to the angular velocity input
(t ) = 200us (t ) + 160(1 e 0.25t ).
9.3.e Plot the input and the responses using Mathcad or
MATLAB.
Problem 9.4 A schematic of a hydraulic system is shown in
Fig.P9.4. The pump, modeled as a pressure source p(t), discharges fluid into a hydraulic circuit consisting of two fluid
resistances, R1 and R 2 , a long run of pipe with inertance, I ,
and a fluid accumulator with capacitance, C.
Table P9.4 Parameter values
R1

R2

Case I

0.5

10

Case II

10

9.4.a Derive the state equations of this system and express


them in vector-matrix form.
9.4.b Calculate the systems eigenvalues for Cases I and II
using the parameter values in Table P9.4. If the system is underdamped, determine the ideal, undamped
natural frequency n, the damped natural frequency
d, and the damping ratio .
9.4.c Use the Laplace transform to solve the state equations
to yield the transfer functions for the state variables
X 2 (s)
X 1 (s)
.
and
U (s)
U (s)
9.4.d Use Laplace transforms and partial fraction expansion, if necessary, to determine the responses
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 mechanical system

Mass M
Lubricating Film
Damping b 2

Pivot

Dashpot b1

Spring K2

Spring K1

L3
L2

Force Source F(t)

Fig. P9.6 Rotational mechanical


system

L1

Compliant Shaft
Spring Constant K
Pulley
Diameter D1

Pulleys Diameters
D2 and D3

Flywheel
Rotational Inertia J1
Flywheel
Inertia J 2

Angular
Velocity
Input (t)

Pulley
Diameter D4

9.4.e Plot the input and the responses using Mathcad or


MATLAB.
Problem 9.5 A translational mechanical system is shown
schematically in Fig.P9.5. The system consists of a force
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
M which slides on a rail with a lubricating film, damping
b2. The parameter values are tabulated. The system is deenergized before the input force act on the system.
Table P9.5 Parameter values
L1

L2

L3

b1

b2

K1

K2

0.5

1,000

1,500

10

9.5.a Derive the state equations of this system.

Fluid Coupling
Damping b

9.5.b Calculate the systems eigenvalues using the parameter values in Table P9.5. If the system is underdamped,
determine the ideal, undamped natural frequency n,
the damped natural frequency d, and the damping
ratio .
9.5.c Use the Laplace transform to solve the state equations
to yield the transfer functions for the state variables
X 1 (s) X 2 (s)
X (s)
,
and 3 .
U (s) U (s)
U (s)
9.5.d Use Laplace transforms and partial fraction expansion, if necessary, to determine the responses of the
state variables to the force input, F (t ) = 400 us (t ).
9.5.e Plot the input and the responses using Mathcad or
MATLAB.
Problem 9.6 A rotational mechanical system is shown schematically in Fig.P9.6. The system consists of an angular
velocity source which acts on the input shaft of a belt drive

9 Transfer Functions, Block Diagrams, and the s-Plane

556
Fig. P9.7 Hybrid electromechanical system

Electric Motor
Transduction

Rotational Inertia J1
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 compliant shaft modeled as torsion spring K, and fluid coupling
with damping b. Note that pulleys D2 and D3 are fixed to the
same shaft. The system is de-energized before the input force
act on the system.
Table P9.6 Parameter values
D1

D2

D3

D4

J1

J2

100

30

9.6.a Derive the state equations of this system.


9.6.b Calculate the systems eigenvalues using the parameter values in Table P9.6. If the system is underdamped,
determine the ideal, undamped natural frequency n,
the damped natural frequency d, and the damping
ratio .
9.6.c Use the Laplace transform to solve the state equations
to yield the transfer functions for the state variables
X (s)
X 1 (s) X 2 (s)
,
and 3 .
U (s)
U (s) U (s)
9.6.d Use Laplace transforms and partial fraction expansion, if necessary, to determine the responses of
the state variables to the angular velocity input,
(t ) = 200 us (t ).
9.6.e Plot the responses using Mathcad or MATLAB.
Problem 9.7 A hybrid electromechanical system is shown
in Fig.P9.7. The input to the system is a voltage which
drives a DC motor through wires with electrical resistance
R. The relationship between the motors current and torque
is TM = iM . Flywheel J1 is attached to the motors shaft and
is supported by bearings with damping b1. Coupled to the
motors shaft is a compliant shaft modeled as torsion spring
K. The compliant shaft drives flywheel J2 which is supported
by bearings with damping b2.

Table P9.7 Parameter values


R

J1

b1

J2

b2

0.6

1,000

0.5

9.7.a Derive the state equations of this system.


9.7.b Calculate the systems eigenvalues using the parameter values in Table P9.7. If the system is underdamped,
determine the ideal, undamped natural frequency n,
the damped natural frequency d, and the damping
ratio .
9.7.c Use the Laplace transform to solve the state equations
to yield the transfer functions for the state variables
X (s)
X 1 (s) X 2 (s)
,
and 3 .
U (s)
U (s) U (s)
9.7.d Use Laplace transforms and partial fraction expansion, if necessary, to determine the responses of the
state variables to the voltage input, v(t ) = 96 us (t )
9.7.e Plot the responses using Mathcad or MATLAB.
Problem 9.8 Draw a block diagram which represents the
state equations and output equations below.
d
dt

x1 a11
x = a
2 21

a12 x1 b1
+
u
0 x2 0

y1 c11 0 x1
y = c

2 21 c22 x2
Problem 9.9 A system equation where u is the input variable
and y is the output variable is given below.
8u = 3

d 3y
d 2y
dy
+
33
+ 114 + 210 y
3
2
dt
dt
dt

Problems

557

9.9.a Express the solution of the system equation as a block


diagram.
9.9.b Derive the state equations from the block diagram
drawn for part a.
9.9.c Solve the state equations for the input u(t) shown in
Fig.P9.9.
9.9.d Plot the output, y(t), using Mathcad or MATLAB.

100
50

u(t) 0
0

100

Problem 9.11 Determine the state equations and output


equations of the system expressed as a block diagram in
Fig.P9.11.
Problem 9.12 For the system shown in the block diagram in
Fig.P9.12.
9.12.a Determine the transfer function Y(s)/U(s).
9.12.b Determine the poles and zeros of the transfer function and plot them on the s-plane.
9.12.c Use Laplace transforms and partial fraction expansion to determine the response of this system to the
step input u (t ) = 4us (t ).
9.12.d Plot the response of the system in Mathcad or MATLAB.

Input

50

u(t) 0
0

0.5

1.0

2.0

1.5

t, sec

-50
-100
Fig. P9.10 Input u(t)

Problem 9.13 A transfer function Y(s)/U(s) is shown as a


block diagram in Fig.P9.13.
Express the transfer function Y(s)/U(s) as a time-domain,
state-space block diagram, using the recursive algorithm of
Sect.9.4.2 to eliminate differentiation of the input variable.
Problem 9.14 A feedforward loop and feedback loop are
shown as block diagrams in Fig.P9.14.
9.14.a Derive the transfer functions, C(s)/R(s). Express the
transfer functions as proper ratios.

2.0

t, sec

-100

9.10.a Express the solution of the system equation as a


block diagram.
9.10.b Derive the state equations from the block diagram
drawn for part a.
9.10.c Solve the state equations for the input u(t) shown in
Fig.P9.10.
9.10.d Plot the output, y(t), using Mathcad or MATLAB.

u2

1.5

Fig. P9.9 Input u(t)

d 2y
dy
d 3y
12u = 2 3 + 29 2 + 273 + 620 y
dt
dt
dt

u1
Input

1.0

-50

Problem 9.10 A system equation where u is the input variable, and y is the output variable is given below.

Fig. P9.11 Block diagram of


state and output equations

0.5

y1

+
+

__
1
s

__
1
s

Output

+
+

+
+

y2
Output

y3
Output

9 Transfer Functions, Block Diagrams, and the s-Plane

558
Fig. P9.12 Block diagram of a
system

1.5

u(t)

0.5s

1
__
s

1
__
s

1
__
s

+
+_ _ _

y(t)

9
28
32

U(s)

s+3
__________________
2s 3 + 18s 2 + 56s + 64

Y(s)

Fig. P9.13 A transfer function

9.14.b Use inverse Laplace transforms to derive the responses


of the systems to the step input, r (t ) = 12us (t ).
9.14.c Plot the responses in Mathcad or MATLAB.
Problem 9.15 A block diagram is shown in Fig.P9.15.
9.15.a Determine the transfer function, C(s)/R(s).
9.15.b Use inverse Laplace transforms and partial fraction
expansion to determine response of the system to the
step input, r (t ) = 18us (t ).
9.15.c Plot the response in Mathcad or MATLAB.
Problem 9.16 A closed-loop feedback system is shown as a
block diagram in Fig.P9.16.
9.16.a
Determine the closed-loop transfer function
C(s)/R(s). Express the transfer functions as proper
ratios.
9.16.b Use inverse Laplace transforms and partial fraction
expansion, if necessary, to derive the response of the
systems to the step input, r (t ) = 12us (t ) .
9.16.c Plot the responses in Mathcad or MATLAB.

Fig. P9.14 a Feedforward loop,


b Feedback loop

Problem 9.17 The block diagram of a closed-loop system is


shown in Fig.P9.17.
9.17.a
Determine the closed-loop transfer function
C(s)/R(s). Express the transfer functions as proper
ratios.
9.17.b Use inverse Laplace transforms and partial fraction
expansion, if necessary, to derive the response of the
systems to the step input, r (t ) = 8us (t ) .
9.17.c Plot the responses in Mathcad or MATLAB.
Problem 9.18 The closed-loop feedback control system,
shown in the block diagram Fig. P9.18, controls the angular velocity of a flywheel, J=2 kg m2, supported on bearings with damping b=0.05N mrad/sec. In controls jargon, a
dynamic system placed under closed-loop control is known
as a plant. The transfer function, GP(s), in the block diagram is the transfer function of the flywheel system, where
the input is the applied torque, T(t), and the output is the angular velocity of the flywheel. Note that the angular velocity
of the flywheel is the output variable, out(s), of the feedback
loop system.
9.18.a Derive the transfer function that relates input force
T(s) to the angular velocity of the flywheel out(s).
The transfer function out(s)/T(s) is the plant transfer function, GP(s), in the block diagram.
9.18.b Determine the closed-loop transfer function which
relates the input, the desired or reference value of
angular velocity of the flywheel, in(s), to the output,

a
R(s)

b
9
_____
s + 12

+
+

C(s)

3
_____
s+7

Fig. P9.15 A cascade of two


blocks

R(s)

9
_____
s + 12

C(s)

3
_____
s+7

s+30
_________
2
s +4s +29

192
______________
2
12s +168s +576

Problems

559

+-

30
_____________
3
4s +20s2 +60s

Flywheel
Inertia J

Hydraulic Motor
High pressure line
Fluid Resistance R

1
____
s+8
Shaft Bearings
Damping b

Fig. P9.16 A closed loop feedback system

the actual or observed angular velocity of the flywheel, out(s).


9.18.c The flywheel was at rest before the system was given
the input, in (t ) = 200 rad/sec us (t ). Determine the
response of the closed-loop system
9.18.d Plot the response of the system using either Mathcad
or MATLAB.

Low pressure return


Fluid power unit
modeled as a
pressure source

Fig. P9.19a The plant to be placed under feedback control

Problem 9.19 The hybrid system shown in Fig.P9.19a is the


plant to be placed under closed loop control per the block
diagram in Fig.P9.19b. The plant consists of a fluid power
unit modeled as a pressure source, p(t), hydraulic lines with
fluid resistance R=5GPasec/m3 on the high pressure line, a
hydraulic motor with displacement vol=24cm3 per revolution and a flywheel J=10kgm2 supported on bearing with
damping b=0.05Nmrad/sec. The output variable is the angular velocity of the flywheel.
9.19.a Derive the plant transfer function which relates the
input pressure P(s) to the angular velocity of the flywheel out(s). If there is a negative (sign inversion)
between the input and output variables, it results
from the linear graph transducer sign convention and
must be removed. A positive input to the plant must
yield a positive output.
9.19.b
Determine the closed-loop transfer function,
out(s)/in(s).

9.19.c Determine the response of the de-energized system


to the input in (t ) = 300 rad/sec us (t ).
9.19.d Plot the response using either Mathcad or MATLAB.
Problem 9.20 An electromechanical system is shown in
Fig. P9.20a. The system consists of a voltage source v(t)
driving a DC motor with electrical resistance R = 3 and
torque constant KT = 7 N m / amp. The motors shaft drives
a flywheel, J = 1.5 kg m 2 , and is coupled to a power screw
with four threads per inch. The nut of the power screw
moves mass M =50kg which slides on a lubricating film
with damping b = 6 N sec/m. The position of mass M is to
be placed under closed-loop, feedback control per the block
diagram in Fig.P9.20b.
9.20.a Derive the plant transfer function which relates the
input, the voltage applied to the DC motor to output,
the velocity of mass M. If there is a negative (sign
inversion) between the input and output variables,

Fig. P9.17 A closed-loop feedback system

+-

4s+6
________________
s3 +11s 2+124s+300
20

Fig. P9.18a Block diagram of


the closed-loop feedback control
system. b The flywheel which
is the plant with its angular
velocity under feedback control

a
in (s)

b
Plant
+

GP(s)

8
7
___
s+2

out (s)

Flywheel
Rotational Inertia J

Bearings
Damping b

Input Torque
T(t)

9 Transfer Functions, Block Diagrams, and the s-Plane

560

in (s)

4(s+10)

P(s)

Plant

GP(s)

The procedure is as follows. First, express the denominator


polynomial in standard form. Clear the highest power of s of
its coefficient

out (s)

1
N (s)
a0
Output ( s ) =
a1
a2
a3
a
s 4 + s3 + s 2 + s + 4
a0
a0
a0
a0

Fig. P9.19b Block diagram of the closed-loop feedback control system

it results from the linear graph transducer sign convention and must be removed. A positive input to the
plant must yield a positive output.
9.20.b
Determine the closed-loop transfer function,
xMass(s)/xin(s).
9.20.c Determine the response of the system to the input,
xin (t ) = 0.5 m us (t ).
9.20.d Plot the response using either Mathcad or MATLAB.

Chapter 9 Appendix

Now, factor the denominator polynomial.


If there is an oscillatory second-order factor, it is generally best to keep it as a second-order polynomial. Consequently, a fourth-order characteristic function yields one of
three possible forms
1
N (s)
a0
a1 3 a 2 2 a 3
a
s4 +
s +
s +
s+ 4
a0
a0
a0
a0

Inverse Laplace Transformation Using Manual


Partial Fraction Expansion

The process of partial fraction expansion requires knowledge of what order polynomial to assume for the numerators of expansion terms, and a fair amount of patience while
expanding polynomials, collecting like terms, and forming a
set of simultaneous algebraic equations, to solve for the coefficients of the numerator polynomials.
To find the time-domain variable, output(t), described by
the Laplace-domain signal, Output(s)

( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )

or
1
N (s)
a0
a1 3 a 2 2 a 3
a
s4 +
s +
s +
s+ 4
a0
a0
a0
a0

N (s)
Output ( s ) =
4
3
a 0 s + a 1 s + a 2 s 2 + a 3 s + a4

Fig. P9.20a The electromechanical plant to be placed under


feedback control

1
N (s)
a0

1
N ( s)
a0

( s + p1 )( s + p2 ) ( s 2 + 2w n s + w n2 )
Flywheel, Rotational Inertia J

DC Motor
Torque constant K T
Electrical Resistance R

Power Screw
Mass M

Lubricating Film
Damping b

Voltage Source v(t)

Fig. P9.20b The block diagram


of the feedback control system

x in (s)

Controller
+

4(s+10)

vMotor(s)

Plant

GP(s)
8

Integrator

vMass(s)

1 x Mass(s)
__
s

Chapter 9 Appendix

561

or
1
N (s)
a0
a1 3 a 2 2 a 3
a
s4 +
s +
s +
s+ 4
a0
a0
a0
a0
=

(s

1
N (s)
a0
2

)(

+ 21w1n s + w12n s 2 + 2 2w 2 n s + w 22n

The next step is to express the Laplace transformed variable


(or signal) as the sum of terms, with the individual factors
determined in step one as denominators. A polynomial is assumed for the numerator, which is one order lower than the
denominator. If the denominator is a first-order polynomial,
then assume a constant for the numerator

set of equations is expressed in matrix notation. The vector


of unknown constants is found by using linear algebra. As
illustrated below, most of the effort in partial fraction expansion lies in expanding polynomials, which can be done
mechanically. A calculator or computer is needed to evaluate
the linear algebra.
The general case will be illustrated, assuming all of the
poles are first order and none are repeated.
1. Factor the denominator in to first- and second-order
polynomials, where the second-order factors have
complex conjugate roots
1
N (s)
a0
a1 3 a 2 2 a 3
a4
s4 +
s +
s +
s+
a0
a0
a0
a0

1
N (s)
a0

( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )
=

A
B
C
D
+
+
+
s + p1 s + p2 s + p3 s + p4

or

A
B
Cs + D
+
+ 2
( s + p1 ) ( s + p2 ) s + 2w n s + w n2

or
1
N (s)
a0
2

)(

+ 21w1n s + w12n s 2 + 2 2w 2 n s + w 22n


=

As + B
s

2. Equate the factored form with the partial fraction expansion, where the numerator terms are polynomials in s of
one order less than the denominator, i.e.,

and

( s + p1 )( s + p2 ) ( s 2 + 2w n s + w n2 )

(s

( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )

D ( s ) = s 2 + 2n s + 2n N ( s ) = As + B

1
N (s)
a0

1
N (s)
a0

+ 21w1n s + w12n

) (

Cs + D
s + 2 2w 2 n s + w 22n

The simplest way to determine the unknown constants, A,


B, C, and D, is to multiply both sides of the equation by the
denominator of the left side, clearing fractions and yielding
a polynomial in the Laplace variable, s. Equating the coefficient terms of like powers of s produces a set of equations
equal in number to the number of unknown constants. The
N (s)
( s + p1 )( s + p2 )( s + p3 )( s + p4 )
a0
( s + p1 )( s + p2 )( s + p3 )( s + p4 )

In this example,
1
N (s)
a0
a1 3 a 2 2 a 3
a4
s4 +
s +
s +
s+
a0
a0
a0
a0

D (s) = s + a N (s) = A

A
B
C
D
+
+
+
s + p1 s + p2 s + p3 s + p4

3. Multiply both sides by the denominator of the left side to


clear fractions. Recall that
D (s) = s4 +

a1
a0

s3 +

a2
a0

s2 +

a3
a0

s+

a4
a0

= ( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )

A
B
C
D
=
+
+
+
( s + p1 )( s + p2 )( s + p3 )( s + p4 )
s
+
p
s
+
p
s
+
p
s
+
p4

1
2
3

9 Transfer Functions, Block Diagrams, and the s-Plane

562

4. Expand the products of the factors into polynomials. This


is the tedious part. It saves time to work one expansion
using generalized variables. Then use the result as a template

N (s) A
=
( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )
a0
s + p1
B
+
( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )
s + p2

( s + a ) ( s + b) ( s + c) = s3 + (a + b + c) s 2
+ (a b + b c + a c) s + (a b c)

C
+
( s + p1 )( s + p2 ) ( s + p3 ) ( s + p4 )
s + p3

N (s)
= A( s + p2 )( s + p3 )( s + p4 ) + B ( s + p1 )( s + p3 )( s + p4 )
a0

D
+
( 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)
= A s 3 + ( p2 + p3 + p4 ) s 2 + ( p2 p3 + p3 p4 + p2 p4 ) s + ( p2 p3 p4 )
a0

+ 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 )
2

+ D s + ( p1 + p2 + p3 ) s + ( p1 p2 + p2 p3 + p1 p3 ) s + ( p1 p2 p3 )
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)
= ( A + B + C + D ) s 3 + A ( p2 + p3 + p4 ) + B ( p1 + p3 + p4 ) + C ( p1 + p2 + p4 ) + D ( p1 + p2 + p3 ) s 2
a0

+ 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 )
a0
+ C ( s + p1 )( s + p2 )( s + p4 ) + D ( s + p1 )( s + p2 )( s + p3 )

6. Create a set of simultaneous equations by equating like


terms of s, where

b3
N ( s ) b0 3 b1 2 b2
s + s + s+
=
.
a0
a0
a0
a0
a0

s3

b0
= A+ B +C + D
a0

s2

b1
= ( p2 + p3 + p4 ) A + ( p1 + p3 + p4 ) B + ( p1 + p2 + p4 )C + ( p1 + p2 + p3 ) D
a0

Chapter 9 Appendix

s1

563

b2
= ( 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

s0

b3
= ( 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
b1
( p2 + p3 + p4 )
a0
= p2 p3 + p3 p4

b2 + p p
2
4
a0
( p2 p3 p4 )
b3
a
0

1
( p1 + p2 + p4 )

1
( p1 + p3 + p4 )

p1 p3 + p3 p4
+ p p

1
4

p1 p2 + p2 p4
+ p p

1
4

( p1 p3 p4 )

( p1 + p2 + p3 ) A
B
p1 p2 + p2 p3

C
+ p p


1
3
D
( p1 p2 p3 )

( p1 p2 p4 )

8. Solve by premultiplying both sides by the inverse of the


coefficient matrix

p +p +p
(
2
3
4)

p2 p3 + p3 p4

+ p2 p4

( p2 p3 p4 )

( p1 + p2 + p4 )

( p1 + p3 + p4 )

p1 p3 + p3 p4
+ p p

1
4

p1 p2 + p2 p4
+ p p

1
4

( p1 p3 p4 )

( p1 p2 p4 )

b0

a
1 0
1

b
( p1 + p2 + p3 ) a 1 A
0
B
p1 p2 + p2 p3 =

b
C
+ p p
2

1
3
D
a
( p1 p2 p3 ) 0
b3

a 0

factor, and then evaluating the equation for the value of s,


which is the pole of that factor

The bad-old-days method can be quicker in some cases, but,


in general, is more involved, and only useful for those who
stubbornly refuse to accept that their beloved middle-school
calculator is obsolete. The method uses algebra rather than
linear algebra. The constants for first-order factors are determined by multiplying both sides of the equation by the

1
N (s)
a0

( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )
=

( s + p1 ) a

N (s)

( s + p1 ) ( s + p2 ) ( s + p3 ) ( s + p4 )

( s + p1 ) A

s = p1

1
N ( p1 )
a0
=A
( p1 + p2 )( p1 + p3 )( p1 + p4 )

s + p1

+
s = p1

( s + p1 ) B
s + p2

+
s = p1

A
B
C
D
+
+
+
s + p1 s + p2 s + p3 s + p4
0

( s + p1 ) C
s + p3

+
s = p1

( s + p1 ) D
s + p4

s = p1

9 Transfer Functions, Block Diagrams, and the s-Plane

564

The process is repeated for to determine the constants, B, C,


and D.
The bad-old-days procedure to determine the constants
for a second-order factor is similar, except it is necessary
to differentiate both sides of the equation with respect to s,
to create two equations, in order to determine the two unknowns.
The special cases for either procedure involve repeated
roots. The general case of repeated roots is not a practical
concern, since repeated roots require two physical subsystems to be identical, or for a second-order factor to be exactly critically damped. Phenomena which require parameters
to have exact values are extremely unlikely.
However, there is one common situation which generates
repeated roots. Such systems have poles at the origin, be1
cause a factor, , in a transfer function represents integras
1
tion. The Laplace transform of a unit step input is also .
s
The Laplace transform of a unit ramp input is the integral of
a unit step. Hence, the unit ramp input is
11 1
=
s s s2
Factors which produce repeated poles at the origin are handled as second-order factors
1
As + B
As + B A s B
1
A B
=

= 2 + 2 2 = + 2
2
2
2
s s
s
s
s
s
s
s
Partial fraction expansion will be illustrated with five examples.
Example One: Use partial fraction expansion to express the
3
second-order transfer function, Output ( s ) = 2
, as
s + 3s + 2
the sum of two first-order transfer functions.
Linear Algebra Method
G (s) =

3
3
A
B
=
=
+
s 2 + 3s + 2 ( s + 1) ( s + 2) ( s + 1) ( s + 2)

3 ( s + 1) ( s + 2)
s 2 + 3s + 2
3 ( s + 1)( s + 2)
s 2 + 3s + 2

A
B
+
=
( s + 1) ( s + 2)
( s + 1) ( s + 2)

A ( s + 1) ( s + 2)
( s + 1)

B( s + 1) ( s + 2)
( s + 2)

s1 0 = A + B
s0 3 = 2 A + B
1 1
2 1

3 = ( A + B) s + (2 A + B)

0 1 1
3 = 2 1

1 1 A
2 1 B

A
3 A
1 1 0
2 1 3 = I B 3 = B




G ( s) =

3
s 2 + 3s + 2

( s + 1) ( s + 2)

Alternative Method
G (s) =

A
B
3
3
=
=
+
s 2 + 3s + 2 ( s + 1)( s + 2) ( s + 1) ( s + 2)
A
B
3
=
+
( s + 1)( s + 2) ( s + 1) ( s + 2)
3( s + 1)
A( s + 1) B ( s +11)
+
=
( s + 1)( s + 2)
( s + 1)
( s + 2)
3 ( s + 1)
( s + 1) ( s + 2)

A ( s + 1)
( s + 1)

B( s + 1)
( s + 2)

B( s + 1)
3
3

=3= A
= A+
( s + 2) s = 1
( s + 2) s = 1
( 1 + 2)
A
B
3
=
+
( s + 1)( s + 2) ( s + 1) ( s + 2)
3( s + 2)
A( s + 2) B( s + 2)
+
=
( s + 2)
( s + 1)( s + 2)
( s + 1)
3 ( s + 2)
( s + 1) ( s + 2)

A( s + 2) B ( s + 2)
+
( s + 1)
( s + 2)

3
3
A( s + 2)
+B
= 3 = B
=
( s + 1) s = 2
( s + 1) s = 2
( 2 + 1)
G (s) =

3 = A( s + 2) + B( s + 1) 3 = ( As + 2 A) + ( Bs + B)

0 1 1 A
3 = 2 1 B

3
3
3
=
+
( s + 1)( s + 2) ( s + 1) ( s + 2)

Example Two: The system shown in the block diagram


1
Fig. A9.1 is given a unit step input, U ( s ) = . Use partial
s

Chapter 9 Appendix

565

Fig. A9.1 Block diagram of system showing a unit step input

3s + 2
_________
s 2 + 7s +12

1
_____
s+2

U(s)
fraction expansion and the Laplace Transform tables to determine the time-domain response.
The s-domain output signal is:

(3s + 2) ( s ( s + 2)( s + 3)( s + 4))

3s + 2
Y (s) =
(9.43)
s s + 2 s 2 + 7 s + 12

)(

The inverse Laplace transform of this signal is not among


our Laplace transform pairs. We will use partial fraction expansion to expand this signal to match the table entries. We
have many possible combinations. Recall the numerator of a
partial expansion term must be a polynomial in s of one order
lower than the denominator. One partial fraction expansion
is
3s + 2

As + B
Cs + D
Y (s) =
=
+ 2
2
s ( s + 2) s + 7 s + 12
s ( s + 2) s + 7 s + 12

Another is
Y (s) =

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)

s+2
s

Cs ( s + 2) ( s + 3) ( s + 4) Ds ( s + 2)( s + 3) ( s + 4)
+

s+3
s+4
3s + 2 = A ( s + 2) ( s + 3) ( s + 4) + Bs ( s + 3) ( s + 4)

+ Cs ( s + 2) ( s + 4) + Ds ( s + 2) ( s + 3)

s ( s + 2) s + 7 s + 12
2

A
B
C
D
+
+
+
s s+2 s+3 s+4

) (
+ C 8s ) + ( Ds

)
+ D6s )

3s + 2 = As 3 + A9 s 2 + A26 s + A24 + Bs 3 + B7 s 2 + B12 s

+ D5s

3s + 2 = ( A + B + C + D ) s 3 + ( A9 + B7 + C 6 + D5) s 2

+ ( A26 + B12 + C 8 + D6) s + A24

s1
s0

0 1 1 1
0 9 7 6
0 = 9 A + 7 B + 6C + 5 D
=
3 26 12 8
3 = 26 A + 12 B + 8C + 6 D

2 24 0 0
2 = 24 A

Because the roots of the second-order factor are real, we will


use the latter.
Linear Algebra Method

Y (s) =

) (

+ C s 3 + 6 s 2 + 8s + D s 3 + 5s 2 + 6 s

s3 0 = A + B + C + D

Y (s) =

) (

3s + 2 = A s 3 + 9 s 2 + 26 s + 24 + B s 3 + 7 s 2 + 12 s

+ Cs + C 6 s
3s + 2

Y(s)

3s + 2

s ( s + 2) s 2 + 7 s + 12

3s + 2
A
B
C
D
= +
+
+
s 4 + 9 s 3 + 26 s 2 + 24 s s s + 2 s + 3 s + 4

(3s + 2) ( s ( s + 2) ( s + 3) ( s + 4))
s 4 + 9 s 3 + 26 s 2 + 24 s
B
C
D
A
= +
+
+
( s ( s + 2) ( s + 3) ( s + 4))
s s + 2 s + 3 s + 4

1 1 1
9 7 6

26 12 8

24 0 0

1
5
6

1 A
5 B
6 C

0 D

0
0

3

2

1 1 1
9 7 6
=
26 12 8

24 0 0

1
5
6

1 1 1
9 7 6

26 12 8

24 0 0

1 A
5 B
6 C

0 D

9 Transfer Functions, Block Diagrams, and the s-Plane

566

0
0
0.042 0 1
0
2 1 0.5 0.25 0 0
=

9 3 1 0.333 3 0

8 2 0.5 0.125 2 0

0
1
0
0

(3) ( 3) + 2
= 2.333 = C
( 3) ( 3 + 2) ( 3 + 4)

0 A
0 B
0 C

1 D

0
0
1
0

Find D. Multiply both sides by ( s + 4), and evaluate for


s = 4.

0.083 A
1 B
=

2.333 C

1.25 D
Y (s) =

3s + 2

s ( s + 2) s 2 + 7 s + 12

( s + 4 ) ( 3s + 2 )
s ( s + 2) ( s + 3) ( s + 4)
+

0.083
1
2.333 1.25
+

+
s
s+2 s+3 s+4

Y (s) =

s ( 3s + 2 )

s=0

sA
sB
+
s s=0 s + 2

+
s=0

sC
s+3

+
s=0

sD
s+4

s=0

+
s = 4

s = 2

( s + 2) B
s+2

+
s = 2

( s + 2) A
s

L {Y ( s )} = 0.083L

s+3

+
s = 2

( s + 2) D
s+4

( s + 3) B
s+2

+
s = 3

s+3

+
s = 3

( s + 3) D
s+4

s+4

s = 4

s + 2
1 2.333
+

s+3

s = 4

1.25

s + 4

1
1 1

+
s + 2
s
1 1
1 1

+1.25

+
s
3
s + 4

Example Three: The system shown in the block diagram


Fig. A9.3 consists of an integrator followed by an overdamped second-order system. The system is given a unit
1
step input of U ( s ) = . Use partial fraction expansion and
s
the Laplace Transform table to determine the time-domain
response.
The s-domain output signal is

1
1
Y ( s) = 2
s ( s + a ) ( s + b )

s = 3

( s + 3) C

( s + 4) D

s = 2

Find C. Multiply both sides by ( s + 3), and evaluate for


s = 3.

y (t ) = 0.083 + e 2t 2.333e 3t + 1.25e 4t


(9.44)

(3) ( 2) + 2
=1= B

2
( ) ( 2 + 3) ( 2 + 4)

( s + 3) (3s + 2)
( s + 3) A
=
s
s ( s + 2) ( s + 3) ( s + 4)
s = 3

2.333

s = 2

( s + 2) C

s+3

0.083
+

s ( s + 2) ( s + 3) ( s + 4)

( s + 4) C

Perform the inverse Laplace transformation to create the


time-domain function.
1

Find B. Multiply both sides by ( s + 2), and evaluate for


s = 2.
=

s = 4

3s + 2
0.083
1
2.333 1.25
=
+

+
s ( s + 2) ( s + 3) ( s + 4)
s
s+2 s+3 s+4

L {Y ( s )} = L

( 2) = 0.083 = A
( 2)(12)

( s + 2 ) ( 3s + 2 )

s = 4

( s + 4) B
s+2

( s + 4) A

(3) ( 4) + 2
= 1.25 = D
( 4) ( 4 + 2) ( 4 + 3)

Alternative Method
Find A. Multiply both sides by s, and evaluate for s = 0.

s ( s + 2) ( s + 3) ( s + 4)

s = 3

Chapter 9 Appendix

567

0.15

Equate like powers of s


s3 :
s2 :
s1 :
s0 :

0.10

y(t)
0.05

0 = B+C+ D
0 = A + 4 B + 3C + D
0 = 4 A + 3B
1 = 3 A.

Express the system of equations in vector-matrix form


0.00

t, sec

0 0
0 1
=
0 4

1 3

Fig. A9.2 Unit step response Eq.9.44

We will assign the values, a = 1 and b = 3. The inverse Laplace transform of this signal is not in our table. Use partial
fraction expansion to expand this signal to match the table
entries
A B
1
C
D
1
= 2+ +
+
Y ( s) = 2

s ( s + a ) ( s + b) s
s s+a s+b

Y ( s) =

s 2 ( s + 1) ( s + 3)

0
1

1
4
3
0

s2
+

s +1

1
3
0
0

0 0
0 1
=
0 4

1 3
1
1
0

1
3
0
0

0 1
0 0
=
0 0

1 0

1
1
0

0
1
0
0

0
0
1
0

0
1

1
4
3
0

1
3
0
0

1 A
1 B
0 C

0 D

0 A A
0 B B
=
0 C C

1 D D

s+3
Conventional Algebra Method (The Hard Way)
Find A. Multiply both sides by s2, and evaluate using s = 0.

)
+ (Cs + Cs 3) + ( Ds

1 = A s 2 + A 4 s + A 3 + Bs 3 + B 4 s 2 + B3s
3

1
4
3
0

0.333 A
0.444 B
=

0.5 C

0.056 D

Ds 2 ( s + 1) ( s + 3)

1 = A( s + 1) ( s + 3) + Bs ( s + 1) ( s + 3) + Cs 2 ( s + 3) + Ds 2 ( s + 1)

+ Ds 2

1
C
D
A B
1
= s2 2 + +
+
s2 2

s
s ( s + 1) ( s + 3)
s s + 1 s + 3

1
1
s2 2
s ( s + 1) ( s + 3)
B
A + s2
s

1 = ( B + C + D ) s 3 + ( A + 4 B + 3C + D ) s 2

+ ( 4 A + 3 B ) s + ( 3 A)

Fig. A9.3 Block diagram of an


overdamped second-order system
with an integrator. The input is a
unit step

1
4
3
0

Cs 2 ( s + 1) ( s + 3)

) (

1
1
0

Bs 2 ( s + 1) ( s + 3)

Expand and collect like powers of s

1
3
0
0

0
1

A s 2 ( s + 1) ( s + 3)

1 A
1 B
0 C

0 D

1
3
0
0

Solve by pre-multiplying both sides of the equation by the


inverse of the coefficient matrix

Linear Algebra Method


First, clear fractions by multiplying both sides by the denominator of the left side.
s 2 ( s + 1) ( s + 3)

1
4
3
0

s
U(s)

1
__
s

0 s=0

=
s=0

C
+ s2
s + 1

0 s=0

D
+ s2
s + 3

1
_________
(s+a)(s+b)

0 s=0

Y(s)

9 Transfer Functions, Block Diagrams, and the s-Plane

568

1
= 0.333 = A
3

2.0

Find B. The technique used to find A does not work to find


1
B. Multiplying by s leaves on the left side, and an s in the
s
denominator of the A term

1.5

y(t) 1.0
0.5

1
1
s 2
s ( s + 1) ( s + 3)

C
B
A
D
= s 2 + s + s
+ s
s + 1
s
s
s + 3

1
1

s ( s + 1)( s + 3)

A

s

s=0

t, sec

( s + 1)

1
1
C
D
A B
= ( s + 1) 2 + +
+

2
s
s s + 1 s + 3
s ( s + 1) ( s + 3)

D
1 1
A
B

= ( s + 1) 2 + ( s + 1) + C + ( s + 1)
2
s + 3
s
s + 3
s
s

s=0

C
+ B + s

( s + 1)

Fig. A9.4 Response function Eq.9.45

This expression cannot be evaluated using s = 0 to find B,


because two terms blow up to infinity

0 s=0

D
+ s

( s + 3)

1 1
2

s + 3
s

0 s=0

The resolution to this dilemma is to multiply both sides by s ,


and then differentiate with respect to s
2

s = 1

A
= ( s + 1) 2
s

0 s = 1

B
+ ( s + 1)
s

D
+ C + ( s + 1)
s + 3

1
C
D
A B
1
= s2 2 + +
+
s 2

s
s ( s + 1) ( s + 3)
s s + 1 s + 3

0 s = 1

0 s = 1

( s + 1) ( s + 3)

= A + Bs +

1 1 1

= = 0.5 = C
2
( 1) 1 + 3 2

Cs 2 Ds 2
+
s +1 s + 3

Find D. Multiply both sides by ( s + 3), and evaluate using


s = 3

d
1
Cs 2 Ds 2
d
= A + Bs +
+

s + 1 s + 3
ds ( s + 1) ( s + 3) ds
d u
using
=
ds v

dv
du
u
ds and ( s + 1)( s + 3) = s 2 + 4 s + 3
ds
v2

1
Cs 2 Ds 2
d
d
A
Bs
=
+
+
+

s + 1 s + 3
ds s 2 + 4 s + 3 ds

(s

( 2 s + 4)
2

+ 4s + 3

= B+

( s + 1) 2Cs Cs
( s + 1)2

( s + 3) 2 Ds Ds
( s + 3)2

(3)

1
1
C
D
A B
= ( s + 3) 2 + +
+

2
s
s s + 1 s + 3
s ( s + 1) ( s + 3)

C
A
B
1 1
+D
= ( s + 3) 2 + ( s + 3) + ( s + 3)
2
s + 1
s
s + 1
s
s

1 1
2

s + 1
s
2

Now evaluate both sides using s=0


(1 + 3)

( s + 3)

4
= 0.444 = B
9

Find C. Multiply both sides by ( s +1) , and evaluate using


s = 1

A
= ( s + 3) 2
s
s = 3

0 s = 3

B
+ ( s + 3)
s

C
+ ( s + 3)
s + a

0 s = 3

+D
0 s = 3

1 1
1
= = 0.056 = D

2
18
( 3) 3 + 1
0.333 0.444 0.5 0.056
1
1
= 2
+

Y ( s) = 2
s ( s + 1) ( s + 3)
s
s +1 s + 3
s

Chapter 9 Appendix

569

Fig. A9.5 Block diagram of


underdamped second-order system with an integrator. The input
is a unit step

The output signal is transformed to the time-domain by the


inverse Laplace transform
y (t ) = L

{Y ( s)} = L

0.333 0.444 0.5 0.056


+

s
s +1 s + 3
s

The Laplace transform is a linear operator, so it has the distributive property


0.444
0.333
1 0.5
y (t ) = L 1 2 L 1

+L
s + 1
s
s
0..056
L 1

s+3
Its linearity allows coefficients to be factored out
1
1
1
y (t ) = 0.333L 1 2 0.444L 1 + 0.5L 1

s
s
s + 1


1
0.056L 1

s + 3
The inverse Laplace transform of these functions of s are
found in the transform pair table


y (t ) = 0.333uramp (t ) 0.444ustep (t ) + 0.5e t 0.056e 3t

(9.45)
Example Four: The system shown in the block diagram
Fig.A9.5 consists of an integrator followed by an oscillatory second-order system. Use partial fraction expansion and
the Laplace transform pairs to determine the time-domain
response.
The algebra in this reduction entails considerably more
effort than in Example Three, because the roots of an oscillatory second-order factor are complex conjugates. Partial
fraction coefficients will be real numbers, except if the input
is a sinusoid. In this example, the input is a unit step. Consequently, A, B, C, and D will be real.
The Laplace-domain output signal is

 n2
1
Y (s) = 2 2
.
2
s s + 2 n s +  n

2n
_____________
s 2+ 2ns + 2n

1
__
s

U(s)

Y(s)

The inverse Laplace transform is not in our transform pair


table. Use partial fraction expansion to expand this to match
1
the table entries. The repeated root of 2 requires two terms,
s
A
B
and . The numerator of the oscillatory second-order fac2
s
s
tor must be a first-order polynomial, Cs + D.
A B
 n2
Cs + D
1
= + +
Y (s) = 2 2
s s + 2 n s + w n2 s 2 s s 2 + 2 n s +  n2
This example will be easier to follow if we assign values to
and  n. Using  = 0.3 and  n = 4,
16
Cs + D
A B
1
= + +
Y ( s) = 2 2
.
s s + 2.4 s + 16 s 2 s s 2 + 2.4 s + 16
This partial fraction expansion is easiest to work using linear
algebra. We will first create a set of simultaneous algebraic
equations for A, B, C, and D, which we will solve using linear algebra. We will then rework the problem finding unknowns individually to illustrate the difficulties created by
1
the repeated root, 2 .
s
Linear Algebra
First, clear fractions by multiplying both sides by the denominator of the left-hand side

16 s 2 s 2 + 2.4 s + 16
2

s s + 2.4 s + 16
+

Bs 2 s 2 + 2.4 s + 16
s

A s 2 s 2 + 2.4 s + 16
s

) + (Cs + D ) s ( s
2

+ 2.4 s + 16

s + 2.4 s + 16

16 = A( s 2 + 2.4 s + 16) + B( s 3 + 2.4 s 2 + 16 s ) + (Cs + D) s 2


16 = ( As 2 + 2.4 As + 16 A) + ( Bs 3 + 2.4 Bs 2 + 16 Bs )
+ (Cs 3 + Ds 2 )
16 = ( B + C ) s 3 + ( A + 2.4 B + D) s 2 + (2.4 A + 16 B) s + 16 A
Form a set of four simultaneous algebraic equations, by
equating coefficients of like powers of s.

9 Transfer Functions, Block Diagrams, and the s-Plane

570

s3 :
s2 :
s1 :
s0 :

16

s + 2.4 s + 16

0 = B+C
0 = A + 2.4 B + D
0 = 2.4 A + 16 B
16 = 16 A

s=0

B
= A + s2
s

Express this system of equations in vector-matrix form

Find B. The technique used to find A does not work to find


1
B. Multiplication by s leaves
on the left side, and an s in
s
the denominator of the A term.

Pre-multiply both sides by the inverse of the coefficient matrix to solve


1
0
0
0

0
1
0

0
0

0

16

1
0
1 2.4
=
2.4 16

16 0
1 1 0
0
1 2.4 0 1

2.4 16 0 0

16 0 0 0


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
This expression cannot be evaluated using s = 0 to find B,
because two terms blow up to infinity

1
0
0
0
1

0
1
0

1
0
1 2.4

2.4 16

16 0

0 1
0 0
=
0 0

16 0

0
1
0
0

0
0
1
0

1
0
0
0

0 A
1 B
0 C

0 D

16 ( 2 s + 2.4)

(s

+ 2.4 s + 16

0 s=0

16
Cs 3 + Ds 2
= A + Bs + 2
s + 2.4 s + 16
s + 2.4 s + 16

Cs + D
B
A
= s2 2 + s2 + s2 2
s + 2.4 s + 16
s
s

s=0

Cs + D
+ B + s 2
s + 2.4 s + 16

16

1
s2 2 2
s s + 2.4 s + 16

A

s

16
Cs + D

1
A B
= s2 2 + + 2
s2 2 2

s s + 2.4 s + 16
s
s s + 2.4 s + 16

Cs 3 + Ds 2
d
16
d
= A + Bs + 2
2
ds
ds s + 2.4 s + 16
s + 2.4 s + 16
du
dv
u
ds
ds and keeping in mind that if K is
v2
dK
a real or complex constant, then
=0
ds
d u
using
=
ds v

d s 2 + 2.4 s + 16

s=0

The resolution to this dilemma is to multiply both sides by s2,


and then differentiate with respect to s

0 A
0 B
0 C

1 D

Conventional Algebra
Find A. Multiply both sides by s2, and evaluate for s = 0

ds
2
s + 2.4 s + 16

16

2
s s + 2.4 s + 16

1
A A
0.15

= I B = B
0.15
C C


0.64
D D

0 16

0 s=0

16
=1= A
16

1 1 0 A
0 0
0 1 2.4 0 1 B
=

0 2.4 16 0 0 C


16 16 0 0 0 D

1
0
1 2.4

2.4 16

16 0

0 s=0

Cs + D

+ s2 2
s + 2.4 s + 16

= B+

(3Cs

)(

+ 2 Ds s 2 + 2.4 s + 16 ( 2 s + 2.4) Cs 3 + Ds 2

(s

+ 2.4 s + 16

s (3Cs + 2 D)( s 2 + 2.4 s + 16) (2 s + 2.4)(Cs 2 + Ds)


16(2 s + 2.4)
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

+ 2.4 s + 16

2
s=0

s (3Cs + 2 D ) s 2 + 2.4 s + 16 ( 2 s + 2.4) Cs 2 + Ds

= B+
2
2
s + 2.4 s + 16

( 2.4)(16)

(16)2

Find C and D. Multiply both sides of the expansion by


s 2 + 2.4 s + 16 , and evaluate twice, using both roots,
s1 = 1.20 + j 3.82 and s2 = 1.20 j 3.82.
Note: It is always best to wait to substitute values for
complex constants, until the last possible point in a reduction, 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.

0 s= 0

= 0.15 = B
Determine C. It is easier to evaluate the expressions, s12 , s22 ,
and ( s1 s2 ) , first. Then substitute them into the expression
for C
s12 = ( 1.20 + j 3.82) = 13.15 j 9.17
2

A s 2 + 2.4 s + 16 B s 2 + 2.4 s + 16 (Cs + D ) s 2 + 2.4 s + 16


16
1
2
2.4
16
+
+
=
+
+
s
s
2 2

s
s
s + 2.4 s + 16
s2
s 2 + 2.4 s + 16

16
s

A s2 + s +
s

s = s1

+
0

B s2 + s +
s

s = s1

)
0

+ Cs s = s + D
1

s = s1

and

A s 2 + 2.4 s + 16
16
=
s 2 s = s2
s2
leading to

+
0

B s 2 + 2.4 s + 16

s = s2

)
0

+ Cs s = s + D
2

s = s2

and
16
16
= Cs1 + D and
= Cs2 + D
2
s1
s22

These two equations with two unknowns yield


16 16
= Cs1 + D (Cs2 + D ) = C ( s1 s2 )
s12 s22
1 1
1
16 2 2
=C
s1 s2 ( s1 s2 )
s2 s2
16 2 22 1 = C
s1 s2 ( s1 s2 )

16
Cs1 = D
and
s12

s22 = ( 1.20 j 3.82) = 13.15 + j 9.17


2

Notice s12 and s22 are complex conjugates


s1 s2 = 1.20 + j 3.82 ( 1.20 j 3.82) = j 7.64
s2 s2
C = 16 2 22 1
s1 s2 ( s1 s2 )
13.15 + j 9.17 ( 13.15 j 9.17 )
C = 16

( 13.15 j 9.17 ) ( 13.15 + j 9.17 ) j 7.64


C = 0.15

9 Transfer Functions, Block Diagrams, and the s-Plane

572

Determine D

f (t ) = ur (t ) = t for t 0

Unit Ramp
1
F ( s ) = s 2

16
16
Cs1 =
( 0.15) ( 1.20 + j 3.82) = D
13.15 j 9.17
s12
In order to add the two terms, multiply the first term by
the ratio of the complex conjugate of the denominator,
13.15 + j 9.17
13.15 + j 9.17 , to create the ratio of a complex number
over a real number. Then divide real and imaginary components of the complex number by the real denominator
13.15 + j 9.17 210.4 + j146.7
16
=
13.15 j 9.17 13.15 + j 9.17
257.0
= 0.819 + j 0.571
D=

16
16
Cs1 =
( 0.15) ( 1.20 + j 3.82)
2
13.15 j 9.17
s1

D = ( 0.82 + j 0.57 ) ( 0.18 + j 57 ) = 0.82 + 0.18


D = 0.64
We now have the coefficients of the partial fraction expansion
16
A B
Cs + D
1

Y (s) = 2 2
= 2+ + 2

s s + 2.4 s + 16 s
s s + 2.4 s + 16
Y ( s) =

Y ( s) =

A
s
1

s2

wn

e w n t sin w n 1 2 t
Oscillatory f (t ) =
2

Unit Impulse
w n2
Response F ( s ) = 2

s + 2w n s + w n2

e w n t sin w n 1 2 t
Oscillatory f (t ) =
1 2
Unit Impulse
s

Response F ( s ) = 2
s + 2w n s + w n2

with

Phase Shift where = tan 1 w d = tan 1 1

The inverse Laplace transformation is a linear operator, allowing us to factor in or out coefficients, as needed. We will need
to factor in and factor out to match the transform table pairs

L
L

L {Y (s)} = y(t )
1

0.15s
0.64

1 0.15
y (t ) = L 1 2
2
2

s
s + 2.4 s + 16 s + 2.4s + 16
s

In order to perform the inverse transformation, we need to


match the two oscillatory terms exactly to the forms of the
Laplace transform pairs
f (t ) = us (t ) = 1 for t > 0

Unit Step
1
F (s) = s

0.15
1
y (t ) = L 1 2 L 1

s
s

0.15s
0.64

1
L 1 2
L 2

s + 2.4 s + 16
s + 2.4 s + 16

B
Cs
D
+
+
s s 2 + 2.4 s + 16 s 2 + 2.4 s + 16

0.15
0.15s
0.64
2
+ 2
s
s + 2.44 s + 16 s + 2.4 s + 16

0.15
= 0.15

0.15s

= 0.15
2
s
2
.
4
s
16
+
+

1

s
s

2
s
2
.
4
s
16
+
+

0.64
0.64
16 1

= L 2
2

16
s
2
.
4
s
16
s
2
.
4
s
16
+
+
+
+

0.64
16
0.64 1

=
2

L 2

16
s
2
.
4
s
16
s
2
.
4
s
16
+
+
+
+

We can now use the Laplace transform pairs with  = 0.3


and  n = 4
1
1
y (t ) = L 1 2 0.15L 1
s
s

s

0.15L 1 2

s + 2.4 s + 16
16

0.04L 1 2

2
4
16
.
s
+
s
+

wn
n
y (t ) = ur (t ) 0.15us (t ) 0.15
e w n t sin w n 1 2 t 0.04
e w n t sin w n 1 2 t
2
1 2

Chapter 9 Appendix

573

2.0

form of the second-order factor we need for the partial fraction expansion:

1.5

30 2 2 s 2 + 20 s + 18
Y (s) =
2

s ( s + 3) s 3s + 7.5s + 24

y(t) 1.0
0.5
0

0.5

1.0

t, sec

1.5

30 2 0.67 s 2 + 6.67 s + 6
Y (s) =
2

s ( s + 3) s s + 2.5s + 8

2.0

s1 , s2 =

Fig. A9.6 Response function, Eq.9.46

2.5

( 2.5)2 + ( 4)(8)

s1 , s2 = 1.25 j 2.54.

y (t ) = ur (t ) 0.15us (t )

4
1
2
2

(0.3)( 4)t
(0.3)( 4)t

0.15
e
e
sin 4 1 ( 0.3) t
sin 4 1 ( 0.3) t 1.27 0.04

2
2

1 ( 0.3)

1 ( 0.3)

y (t ) = t 0.15 + 0.16e 1.2t sin (3.82t 1.27 ) 0.04 0.17e 1.2t sin (3.82t )

Example Five: Laplace transforms, transfer functions, and


partial fraction expansion are the tools, which allow us to
determine the response of systems to arbitrary inputs without resorting to a numerical solver, such as the RungeKutta
method. Our only restriction is that the inputs we apply to
the systems transfer function have a Laplace transform. (It
is possible to approximate almost any function as a sum of
Laplace transforms, but then we may as well use a numerical
solver.) We will use a first-order stable exponential growth as
the input to a system. This is a common input, which occurs
when the dynamic response of the power supply prevents
us from reasonably approximating the input as a Heaviside
step input.
The system shown in the block diagram Fig.A9.7 consists of an integrator, followed by an oscillatory second-order system. Use partial fraction expansion and the Laplace
Transform tables to determine the time-domain response.
First, note the second-order transfer function is not in
standard form. Put it into standard form. Then calculate the
eigenvalues of the characteristic function to determine the

Fig. A9.7 Response function


Eq.9.45

(9.46)

Complex conjugate eigenvalues. Therefore, keep the characteristic function as a second-order polynomial
30 2 0.67 s 2 + 6.67 s + 6
Y (s) =

s ( s + 3) s s 2 + 2.5s + 8
Y (s) =

40 s 2 + 400 s + 360

s 2 ( s + 3) s 2 + 2.5s + 8

Note the s 2 factor in the denominator. Equate the signal to its


partial fraction expansion with unknown numerator constants
Y (s) =
Y (s) =

A
s

40 s 2 + 400 s + 360

s 2 ( s + 3) s 2 + 2.5s + 8
+

B
C
Ds + E
+
+ 2
s s + 3 s + 2.5s + 8

Multiply both sides by the denominator of the left side. Cancel common factors

30
______
s(s + 3)
U(s)

2
__
s

2s 2 + 20s +18
____________
3s2 + 7.5s +12

Y(s)

9 Transfer Functions, Block Diagrams, and the s-Plane

574

( 40s

)
( s + 3) ( s

+ 400 s + 360 s 2 ( s + 3) s 2 + 2.5s + 8


s2

A
s

+ 2.5s + 8

Express in matrix form

s 2 ( s + 3) s 2 + 2.5s + 8

B 2
s ( s + 3) s 2 + 2.5s + 8
s
C 2
s ( s + 3) s 2 + 2.5s + 8
+
s+3
Ds + E
s 2 ( s + 3) s 2 + 2.5s + 8
+ 2
s + 2.5s + 8

40 s 2 + 400 s + 360 = A ( s + 3) s 2 + 2.5s + 8

+ Bs ( s + 3) s + 2.5s + 8

+ Cs 2 s 2 + 2.5s + 8

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

s + 2.5s + 8s + 3s + 7.5s + 24 s
+ B

s 4 + 5.5s 3 + 15.5s 2 + 24 s

(
+ D (s
+ E (s

+ C s 4 + 2.5s 3 + 8s 2
4

+ 3s
+ 3s

)
)

40 s 2 + 400 s + 360 = A s 3 + 5.5s 2 + 15.5s + 24

(
+ C (s
+ D (s
+ E (s

+ B s + 5.5s + 15.5s + 24 s
+ 2.5s + 8s

+ 3s

+ 3s 2

)
)

Collect like powers of s


s4 :
s3 :
s2 :
s1 :
s0 :

0 = B+C + D
0 = A + 5.5B + 2.5C + 3D + E
40 = 5.5 A + 15.5 B + 8C + 3E
400 = 15.5 A + 24 B
360 = 24 A.

1
1 1 0 0
0
1
5.5
2.5
3 1 0

5.5 15.5 6 0 3 40

0 0 0 400
15.5 24
24
0
0 0 0 360
1

Expand polynomials

Solve

1
1 1 0 0
1
1 1 0 A
0
1

5.5 2.5 3 1 1
5.5 2.5 3 1 B

= 5.5 15.5 8 0 3 5.5 15.5 8 0 3 C



0 0 0 15.5 24
0 0 0 D
15.5 24
24
0
0 0 0 24
0
0 0 0 E

+ ( Ds + E ) s 2 ( s + 3)

1
1 1 0 A
0 0
0 1
5.5 2.5 3 1 B


40 = 5.5 15.5 8 0 3 C



0 0 0 D
400 15.5 24
360 24
0
0 0 0 E

1
1
0
1
5.5 2.5

5.5 15.5 8

0
15.5 24
24
0
0

1
3
0
0
0

0
1
3

0
0

0 A
0 B

40 = C

400 D
360 E

A 15
B 6.98


C = 5.61


D 1.37
E 35.3

Substitute into the partial fraction expansion


Y (s) =
Y (s) =

A B
C
Ds + E
+ +
+ 2
2
s s + 3 s + 2.5s + 8
s

15 6.98 5.61 1.37 s + 35.3


+

s
s + 3 s 2 + 2.5s + 8
s2

The inverse Laplace transforms must be an exact match to


the transform pairs. The relevant pairs are:
f (t ) = us (t ) = 1 for t > 0

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

y(t) 20

f (t ) = e at

Unit Exponential Decay


1
F (s) =
s+a

10
0

wn

e w n t sin w n 1 2 t
Oscillatory f (t ) =
2

Unit Impulse
w n2
Response F ( s ) = 2

s + 2w n s + w n2

)
)

The first step is to express the final term as the sum of two
terms, rather than a sum over a common denominator
15 6.98 5.61 1.37 s + 35.3
+

s
s + 3 s 2 + 2.5s + 8
s2
1.37 s
35.3
15 6.98 5.61
Y (s) = 2 +

2
2
s
s + 3 s + 2.5s + 8 s + 2.5s + 8
s
Y (s) =

The first three terms will have the proper form, after the constants in the numerators are factored out
y1 (t ) + y2 (t ) + y3 (t )
=

= 15

15
1 6.98
1 5.61

2+

s
s
s + 3
1 1
1 1
1 1

5.61
2 + 6.98
s + 3
s
s

y1 (t ) + y2 (t ) + y3 (t ) = 15 ur (t ) + 6.98 us (t ) 5.61 e 3t
The time-domain functions corresponding to the two second-order oscillatory factors are expressed in terms of the
natural frequency, n, and damping ratio, . We have already
calculated the real exponential decay factor and the damped
natural frequency, since these correspond to the real and
imaginary terms of the eigenvalues
s =  + j =  n  n 1  2

0.5

1.0

1.5

t, sec

2.0

Fig. A9.8 Response function Eq.9.47

e w n t sin w n 1 2 t
Oscillatory f (t ) =
2
1
Unit Impulse
s

Response F ( s ) = 2
2
+
2
s
w
n s + wn

with

Phase Shift where = tan 1 w d = tan 1 1

Unfortunately, there is a real coefficient in the time-domain


functions, which is written in terms of n and . We could
solve for them using eigenvalues, but it is easier to calculate
them by equating like coefficients of the denominator, and
the standard form of the Laplace transform table
2

2
n

s + 2 n s +  = s + 2.5s + 8

 n = 8 = 2.83

=

2.5
= 0.44
2 n

Comparing the two oscillatory terms to their corresponding


Laplace transform table entries, we see that we can match the
first term by factoring out its numerator coefficient
1.37 s

L 1 s 2 + 2.5s + 8

y4 ( t ) =

L 1 s 2 + 2.5s + 8

y4 (t ) = 1.37

The second oscillatory term requires  n2 in the numerator to


match the table entry. We provide needed value by multiplying both the numerator and denominator by  n2 . This allows
us to factor out the constant, 35.3, and factor in  n2 with no
net change
y5 (t ) =

8
y5 (t ) =
8

35.3
y5 (t ) =
8

35.3

2
s + 2.5s + 8
1

35.3

2
s + 2.5s + 8
1

2
s + 2.5s + 8

9 Transfer Functions, Block Diagrams, and the s-Plane

576

We are now ready to evaluate the inverse Laplace transforms


of the oscillatory terms
y4 (t ) + y5 (t ) = 1.37

2
s
+
2
.
5
s
+
8

8
35.3

8
s + 2.5s + 8

Note there is no unit ramp function in Mathcad. The unit


ramp, for time, t > 0, is t. Multiplying t by the unit step
function prevents it from being evaluated for negative time.
Using Mathcads notation, (t) for the Heaviside unit step
function, the unit ramp function is
ur (t ) = t (t )

y4 (t ) + y5 (t )
= 1.37

2.83
1 ( 0.44)

35.3

e 1.25t sin ( 2.54t )


1

1 ( 0.44)

References and Suggested Reading

e 1.25t sin ( 2.54t )


2.54
where  = tan 1 d = tan 1
= 1.11
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 )

3t
1.25t
sin ( 2.54t )
 y (t ) = 15ur (t ) + 6.98us (t ) 5.61e 4.32e

+ 4.91e 1.25t sin ( 2.54t 1.11)

(9.47)

Ogata K (2003) System Dynamics, 4th edn. Prentice-Hall, Englewood


Cliffs
Ogata K (2009) Modern Control Engineering, 5th edn. Prentice-Hall,
Englewood Cliffs
Rowell D, Wormley DN (1997) System dynamics: an introduction.
Prentice-Hall, Upper Saddle River

10

Frequency Response

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 systems 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 sinusoids magnitude or amplitude differs from that of
the input sinusoid. Input and output sinusoids peak at different times. The steady-state response of a system to sinusoidal inputs across a range of frequencies is called the systems
frequency response, which is represented graphically by Bode and Nyquist plots.

10.1Overview of Sinusoidal Excitation


and Frequency Response
In the previous chapters, we solved first- and second-order
linear differential system equations for step inputs. We used
those superposition step inputs and step responses to create
pulse inputs and their corresponding pulse responses. We observed if pulse inputs are reduced in duration, then any arbitrary input function can be approximated, by superimposing
time shifted pulses and the systems response, and calculated
by superimposing the corresponding pulse responses.
Sinusoids are the second fundamental type of input we must
investigate. Cyclic or periodically varying excitation of physical
systems occurs naturally by wind, waves, and tides. Mechanical occurrences can be attributed to rotating and reciprocating
machine elements. Electrical occurrences arise from alternating
electrical current used to power electrical machines and circuits.
These inputs are either sinusoidal, or they may be reasonably
modeled as sinusoidal. Further, just as scaled and time shifted
steps or pulses are superimposed to create an arbitrary nonperiodic input, scaled and phase-shifted sinusoids of different frequencies can be superimposed to create an arbitrary period input.
We will use the force source mass-damper system and the
force source spring-mass-damper system from Chap.3 as
the example systems in our investigation of the response of
first and second-order energetic systems to sinusoidal inputs.

The systems are shown in Figs.3.13 and 3.31. The system


equations for the velocity of the mass in those systems are
Eqs.3.26 and 3.34. All are reproduced below for reference.


F (t ) M dv1g
F (t ) dv1g b
=
+ v1g
=
+ v1g
b
b dt
M
dt
M
2
1 dF (t ) d v1g b dv1g K
=
+
+ v1g
M dt
M dt
M
dt 2

(3.26)
(3.34)

Forming the transfer functions, starting with the massdamper system:


F (t )
dv1g
b

=L
+ L v1g
M
dt
M

dv
1
L { F (t )} = L 1g + b L v1g
M
dt M

{ }

1
b
F ( s ) = sV1g ( s ) + V1g ( s )
M
M
1
b

F ( s ) = s + V1g ( s )

M
M


V1g ( s )
F (s)

1
M

b
s+
M

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_10, Springer Science+Business Media New York 2014

(10.1)

577

10 Frequency Response

578
Fig. 3.13a Translational mechanical mass-damper system
driven by a force source,
b Linear graph of the massdamper system driven by a force
source

x,v
Mass
M

F(t)

Lubricating
fluid film
Damping b

F(t)

g
Fig. 3.31a Force source-springmass-damper system of Fig.3.23,
b Linear graph, Fig.3.24b

g
Lubricating fluid
Damping b
Now, the spring-mass-damper system:
2
d v1g b dv1g K

+ v1g
2 +
M dt
M
dt

2
dv1g K
d v1g b
dF (t )

= L 2 + L
+ L v1g
dt
dt M
dt M
1
b
K
sF ( s ) = s 2V1g ( s ) +
sV1g ( s ) + V1g ( s )
M
M
M
1
b
K

sF ( s ) = s 2 +
s + V1g ( s )

M
M
M

1
L
M

1 dF (t )

=
M dt

x,v

F(t)

{ }

1
s
M
=
b
K
F (s)
s2 +
s+
M
M

V1g ( s )

(10.2)

Evaluation of the transfer functions, Eqs.1.1 and 1.2, for the


parameters, M = 2 kg, b = 2 N sec/m, and K=200 N/m,
yields:
1
V1g ( s )
V1g ( s ) 0.5
M
=

=

b
F (s)
F (s) s + 1
(10.3)
s+
M
and
1
s
V1g ( s )
0.5s
M
=

= 2
b
K
 F (s)
F ( s ) s + s + 100
s2 +
s+
M
M
V1g ( s )

(10.4)

F(t)
g

The step responses of the initially de-energized massdamper and the spring-mass-damper systems to a step input
of F(t)=10 N us(t) are shown in Fig.10.1. The response of
the initially de-energized, first-order, mass-damper system
to the sinusoidal input, F (t ) = 10 N sin(w t ) for the frequencies of 2rad/sec and 10rad/sec, are shown in Fig.10.2. Figure 10.3 shows the response of the initially de-energized,
second-order, spring-mass-damper system to the same sinusoidal force inputs.
Inspection of the responses of the first-order, mass-damper system and the second-order, spring-mass-damper system
to the same magnitude force as both step input and as a sinusoidal input, with the latter at the frequencies of =2 and
=10rad/sec, yields the following observations.
The transient period of each system appears to be independent of the sinusoidal excitation frequency. It is of the
same duration as the transient response of the systems to
the step input, Figs.10.1a and 10.2a and b, and Figs.10.1b
and 10.3a and b. The transient period is approximately
8 to 10sec.
The peak values of the sinusoidal response of the firstorder, mass-damper system decreased during the transient period, with the mean value of the response equal to
zero in steady-state.
The amplitude of the sinusoidal response of the first-order,
mass-damper system decreased, when the input force frequency was increased from =2 rad/sec to =10 rad/
sec, Figs.10.2a and b.
The amplitude of the sinusoidal response of the secondorder, spring-mass-damper system increased, when the

10.1 Overview of Sinusoidal Excitation and Frequency Response


Fig. 10.1 Responses of the
initially de-energized a firstorder mass-damper system, and
b the second-order spring-massdamper system to the step input
of 10N

M = 2 b = 2 F0 = 10

v1g
m
___
sec

579

3
2

v1g

10

t, sec

12

-0.5

14

rad
___
M = 2 b = 2 F0= 10 = 2 sec

t, sec

10

12

14

rad
M = 2 b = 2 F0= 10 = 10 ___
sec

0.5

m
___
sec

-2

1.0

v1g

-4

m
___
sec

Fig. 10.3 Responses of the initially de-energized second-order


mass-spring-damper system to a
sinusoidal force of 10N with a
the frequency of 2rad/sec, and b
the frequency of 10rad/sec

v1g

-0.25

Fig. 10.2 Responses of the initially


de-energized first-order mass-damper
system to a sinusoidal force of 10N
with a the frequency of 2rad/sec, and
b the frequency of 10rad/sec

M = 2 b = 2 K = 200 F0 = 10

0.25

v1g
m
___
sec

0.5

-0.5

0.2

10

t, sec

12

rad
M = 2 b = 2 K = 200 F0= 10 = 2 ___
sec

0.1

b
v1g

t, sec

10

12

14

rad
___
M = 2 b = 2 K = 200 F0= 10 = 10 sec

m 0
___
sec

m 0
___
sec

-4

-0.1
-0.2

-1.0

14

input force frequency was increased from =2 rad/sec to


=10 rad/sec, Figs. 10.3a and b.
An exponential decay envelope bounds the first-order,
mass-damper systems responses at both input frequencies and the second-order, spring-mass-damper systems
response to the input force at =2 rad/sec, Figs. 10.2a
and b and 10.3a.
An exponential growth envelope bounds the secondorder, spring-mass-dampers response to the input force
at =10 rad/sec, Fig. 10.3b.
The transient oscillations of the second-order, springmass-damper systems step response and sinusoidal
response at =2rad/sec appear to be the same frequency,
Figs.10.1b and 10.3a. The latter transient oscillations are
superposed on the larger sinusoid, which remains in the
systems steady-state.
The steady-state portions from 10 to 14second of the sinusoid responses of the mass-damper system are plotted as

t, sec

10

12

14

-8

t, sec

10

12

14

Fig.10.4, and of the spring-mass-damper system as Fig.10.5.


Added to these plots is the sinusoidal input force, scaled on
the right side.
Inspection of the steady-state plots of the input force and
the output velocity yields additional observations.
The first-order, mass-damper systems output, the velocity
of the mass, peaks after (later than) the force input at both
=2 and =10rad/sec, Figs.10.4a and b.
The second-order, spring-mass-damper systems velocity
output peaks after (later than) the force input at =2 but
simultaneously at =10 rad/sec, Figs. 10.5a and b.
All of the observations above can be understood through investigation of the transient portion of the sinusoidal response
of the two systems. We shall find that, as with step inputs,
the sinusoid response of first-order systems has less variation than the response of second-order systems. We shall also
find that working with sinusoids in the time-domain requires
a great deal of effort which can be eliminated, when we are

10 Frequency Response

580
Fig. 10.4 Steady-state portion
of the responses of the first-order
mass-damper system to a sinusoidal force of 10N with a the
frequency of 2rad/sec, and b the
frequency of 10rad/sec

rad
___
M = 2 b = 2 F0 = 10 = 2 sec

F(t)

v1g

v1g

0.2

0.6

0.3

11

12

t, sec

13

-12
14

0.1

F(t)
v1g

m 0
___
sec
-0.1

-0.2
10

11

12

t, sec

only interested in the steady-state response, by working in


Laplace-domain, using what is termed the Frequency Response Equation, which we will develop in Sect.10.2.
Frequency response refers to the steady-state response
of a system to a sinusoidal input. The system reaches steadystate, when the natural (homogeneous or unforced) portion
of the response has decayed to insignificance after four or
five time constants, leaving just the forced (particular, excited, or driven) response. Steady-state has the same meaning,
whether the input to the system is a sinusoid or a step. The
two steady-state responses, however, are very different. Since
the forced (particular) response of a linear system must have
the same form as the input, when a system subjected to a step
input is in steady-state, every power variable in the system
is constant. In contrast, when the same system subjected to
a sinusoidal input reaches steady-state, every power variable
in the system is a sinusoid. Likewise, when a step response
reaches steady-state, we see that there is no further change
with time. When the same system reaches steady-state in response to a sinusoidal input, we see that each cycle is identical to the cycles which preceded it and which will follow it.
If we were concerned that the initial application of a sinusoidal load to a linear model would create a critical condition for machine design, we would need to investigate the
entire duration of the transient response. We can formulate
the transient response of a system to a sinusoidal input, either
by using transfer functions and Laplace transforms, or a set
of first-order differential equations in state-space. In either

13

-0.6

rad
M = 2 b = 2 K = 200 F0 = 10 = 2 ___
sec

v1g

F(t)

12

v1g

6
0

F(t)
N

-6

-0.3

-6

4
10

12

rad
M = 2 b = 2 F0= 10 = 10 ___
sec

F(t) v1g
0
m 0
N ___
sec

m 0
___
sec

Fig. 10.5 Steady-state portion of


the responses of the second-order
mass-spring-damper system to a
sinusoidal force of 10N with a
the frequency of 2rad/sec, and b
the frequency of 10rad/sec

10

11

12

t, sec

13

14

-12

rad
___
M = 2 b = 2 K = 200 F0= 10 = 10 sec

12

12

v1g

F(t)

F(t) v1g
0
m 0
N ___
sec

-6

-4

-6

-12
14

-8

F(t)

10

11

12

t, sec

13

14

-12

case, we would also use a computer to calculate and plot


the time history. In contrast, if we were concerned that the
steady-state response of a linear model to a sinusoidal input
contained the critical condition, we could easily calculate the
magnitude and phase shift manually, obviating any need to
plot the steady-state response to find the critical value.
From our physical experience, we expect the response of
a physical system to a sinusoidal input to be periodic, but
not necessarily sinusoidal. Further, we expect the magnitude,
frequency, and phase shift of the periodic output to depend
on (1) the amplitude of input, (2) the frequency of the excitation, and, in some way, on (3) the nature of the system.
When we work with linear models of physical systems, our
expectations can be more definite. The response of a linear
model to a sinusoidal input will be sinusoidal. Further, the
amplitude of the input sinusoid scales the amplitude of the
output variable as a direct proportion. We must determine
how the frequency of the input affects the amplitude of the
output, as well as the phase relationship between the input
and output sinusoids.

10.1.1Calculating Magnitude and Phase Angle


from a Response
Physical experience leads us to expect that the magnitude of
the output variable of a system driven by a sinusoidal input
depends on three variables: (1) the nature of the physical

10.1 Overview of Sinusoidal Excitation and Frequency Response

max in

581
Period T

Input

2A

2A|G(j)|
A|G(j)|
Output

Shift t

max out
0

min out

Input

min in

Output

Fig. 10.6 Oscilloscope trace measurements used to calculate the magnitude ratio. The factor of two cancels in the ratio of output/input

Fig. 10.7 Oscilloscope trace measurements used to calculate period


T and frequency , of the two signals and phase shift between them

system, (2) the amplitude of the input sinusoid, F0 in our example system, and (3) the frequency of the input, . Further,
we expect the output to have the same frequency as the input,
when the system is steady-state. We do not expect the two
sinusoids to necessarily peak at the same time. One may be
shifted, relative to the other. We have seen these phenomena
in Figs.10.2 through 10.5.
The magnitude ratio is the ratio of the amplitude of the
output sinusoid over the amplitude of the input sinusoid. The
factor, which represents the effect of the system on the magnitude of the output variable, equals the magnitude of the
transfer function evaluated as a convention function, where
the argument is the input frequency multiplied by the complex number, j, |G(j)|, as we shall see when we develop the
Frequency Response Equation If we had measurements of
the input and output sinusoids on a two-channel oscilloscope
or data acquisition system, we can easily calculate the magnitude ratio and, thus, |G(j)|.
The time difference between the peaks of the input and
output variables is normalized by the period of the cycle, T,
to yield the phase angle, . In other words, the phase angle
is the fraction of a cycle by which the input and output
sinusoids are shifted in time, with the fraction of a cycle (or
circle) expressed as an angle.
The magnitude ratio is calculated from traces of the
input and output sinusoids, by measuring the peak to
peak amplitudes, rather than the mathematical amplitude,
A, Input (t ) = A sin(w t ), which is from zero to the peak,
Fig.10.6. The rationale for measuring twice the amplitude is
the ease of the measurement.

and rad/sec may be presented as 1/sec. The symbol, f, is used


for frequency in Hz. The symbol, , represents angular frequency in rad/sec.
If there is a phase difference between the input and output signals, measure the smaller time difference, t, between
peaks of the two signals. It is impossible to determine from
a single pair of inputoutput traces, whether the time shift
between input and output is the larger or smaller of the two
possible time shifts. In practice, one determines which is the
true time shift, by starting with a low input frequency, and
observing the shift as the frequency is increased. Alternatively, if the transfer function of the system is available, phase
angle can be calculated, as we shall see in Sect.10.3.

Magnitude Ratio =

output (t ) peak to peak


input (t ) peak to peak

2 A G ( jw )
= G ( jw )
2A

The duration of sinusoid cycle is its period, T. The periods


of the input and output sinusoid are equal in steady-state. Instruments report data and are set, by using frequency in units
of Hz, or cycles/sec. However, system dynamics calculation
must be made by using frequency in units of rad/sec. Beware
that all angular measures are dimensionless, so that both Hz

t
T

= phase shift as a fraction of a cycle


2 rad t
= G ( jw )
cycle T

The ratio, t /T , is the phase shift as a fraction of a cycle.


Multiplying the phase shift as a fraction of a cycle (or circle)
by 2 rad yields the phase shift in radians, Fig.10.7.
2 rad t
= G ( jw ) in radians
cycle T
Likewise, to calculate the phase angle in degrees, multiply
the fraction of a cycle by 360. We will routinely use degrees
for geometric constructions, since we can easily visualize an
angle in degrees. We will routinely use radians for calculations, since radians are the natural angular measures.

t
T

360o t
= G ( jw ) in degrees
cycle T
2 rad
T period =
rad
w
sec
2 rad T time shift and phase shift

10 Frequency Response

582
Transient

Positive Impulse

Steady-State

F0

v1g
m
___
sec

F0 sin(0t) 0
0

-F0
-2

=4

=2
0

t, sec

=8
10

15

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 8rad/sec

Unfortunately, the time shift and period data often contain


measurement error, due to the difficulty of identifying identical 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
the cycle needed for the measurements. In practice, electronic noise is superposed on the signals, limiting the accuracy
of peak-finding algorithms. The difficulty is typically compounded by a degree of non-linearity, which deforms the output sinusoid. The human eye is often a more accurate judge
of the peak value or midpoint of a fuzzy trace than are algorithms. Filtering data to remove noise must be done with care.
As we will see later, a noise filter imparts its own dynamic
characteristics to the filtered signal. The safest noise reduction method provides a simple running average that eliminates outlying values, which an algorithm may misidentify as
peaks, without introducing dynamic characteristics.

10.1.2Transient and Steady-State Response


of a First-Order System
The parameter values of M=2 and b=2 were used to create
the step response plot in 10.1a and the sinusoidal response
plots of Figs.10.2 and 10.4. The eigenvalue and the time
constant of the system is one second, per Eq.1.3. The duration of the transient response is defined by the engineer as
a multiple of the time constant. We will use five time constants as the end of the transient period, which corresponds
to 99.3% progression to steady-state.
Plots of the output variable, the velocity of mass
v1g (t ), for three different input force frequencies,
w = 2 rad/sec , 4 rad/sec, and 8 rad/sec, Fig.10.8, show the
amplitude of v1g (t ) decreasing with increasing frequency
of the input force, F0 sin(w t ). This is always the case, if
the system is first order, and the input is not differentiated.
The response plots show a sag of the peak values during

__
T
__
0 = 2

2
__
0 = T

t, periods

3
__
0

Fig. 10.9 Sinusoidal force, F (t ) = F0 sin(w t ). Cross-hatching indicates the half cycle of positive impulse

Positive Impulse
F0

F0 sin(20t) 0
-F0
0

___
__
20 0

2
__
0

3
__
0

t, periods

Fig. 10.10 Doubling the frequency of the input force to 2w 0 halves the
duration of the positive impulse

the transient period, with the responses centered on zero in


steady-state.

10.1.2.1 Magnitude as a Function of Frequency


The decrease in amplitude of the output with increased frequency of the input can be understood as an instance of impulsemomentum theorem. Recall impulse is the time integral of force, Fig.10.9. Increasing the frequency of the input
force decreases the half period of the sinusoid. The positive
impulse delivered to the system, before the force reverses
sign, is reduced as the frequency is increased, Fig.10.10,
thereby reducing the maximum positive velocity. The same
analysis applies to the negative impulse and the maximum
negative velocity.
We must also consider the damping in this system. We
must deduct the portion of the applied force, which is balanced by the shear of the fluid film, to yield the force acting
to accelerate the mass, FM. The force acting to accelerate the
mass is plotted with input force F(t) and the output, the velocity of the mass, v1g, in Fig.10.11. The positive and negative
impulse of FM is cross-hatched. Also shown is the time shift
t between input force F(t) and output mass velocity, v1g.

10.1 Overview of Sinusoidal Excitation and Frequency Response

Transient

Steady-State

F(t)

v1g

m
___
sec

-3

-5

FM
5

= 0.5
10

Transient

15

F(t)

10

v1g

FM

-5

=1

-6
0

t, sec
5

Transient

v1g

5
0

-3

-5

FM

-6
0

=2
2.5

10

F(t)

FM

Steady-State

m
___
sec

F(t)
N

-10
15

10

v1g

F(t)
N
FM
N

-10
5

t, sec

7.5

x 1g 2
m

1
0

=2
0

10

15

t, sec

20

25

Fig. 10.12 Displacement of the mass, x1g, for an input force frequency
of =2rad/sec, calculated by trapezoidal integration of the velocity, v1g

-3

FM

Steady-State

m
___
sec

F(t)
N

-10
25

20

t, sec

10

v1g

-6

v1g

583

10

Fig. 10.11 v1g, F(t), and the force acting to accelerate the mass, FM,
are plotted. The cross-hatch area is the impulse accelerating the mass
during the transient period. a Input force frequency, w = 0.5rad/sec.
Time shift, t=0.93sec. b Input force frequency, w = 1rad/sec. Time
shift, t=0.78sec. c Input force frequency, w = 2 rad/sec. Time shift,
t=0.56sec

The migration of the mean value of the velocity of mass,


v1g, from its initial positive mean to a mean of zero, can be
understood by comparison with the force acting to accelerate
mass, FM, as the difference between applied sinusoidal force
F(t) and the damping force acting to shear the lubricating
fluid film, Fb.
FM = F (t ) Fb FM = F (t ) bv1g

With reference to Figs.10.11b and c, the positive impulse


of FM, that is, the area under the first peak, ends when FM
swings negative, while the input force is still positive, because the mass has gained sufficient velocity for the damping force to exceed the input force at that moment. The following negative impulse of FM is larger than the initial positive impulse, decelerating the mass. That, in turn, leads to
a larger positive impulse of FM, because the velocity swings
positive, later in that half cycle. At the end of the transient
period, the positive and negative impulses of FM are equal,
with the velocity of the mass centered at zero. Do not expect the displacement of the mass to be centered about zero.
The initial asymmetry of the mass velocity leads to a nonzero mean displacement, as shown in Fig.10.11c, for an
input frequency of =2 rad/sec.

10.1.2.2 Phase Angle as a Function of Frequency


The relationship between input frequency and the phase shift
and the input and output variables is more difficult to visualize or calculate. We have an intuitive awareness that the
magnitude of a vibration or oscillation is dependent on the
frequency of the input, because we can feel or hear the speed
of an engine, a vehicle, or of rotating machinery with its resulting vibration. We do not have a similar intuitive sense for
phase shift. We cannot demarcate the time difference in peak
amplitudes of the input and output, as a fraction of the period
of the oscillation.
Phase shift data reveal a great deal about a dynamic system. In practice, however, phase shift data tend to be less
accurate than amplitude data. When we view the input and
output sinusoids on a two-channel oscilloscope, measurement of the time shift between the two wave forms requires
human judgment, which leads to error. Use of automatic
measurement does not improve the accuracy of phase-shift
data. Peak values identified by an instruments algorithm are
often noise. In any event, judicious parsing to remove spurious data is sufficient to yield useful results.

10 Frequency Response

584

The steady-state phase angle, , of the mass-damper system is calculated below, by using the time shift data reported
in Fig.10.11. The negative sign indicates the velocity of the
mass peaks after, or lags, the input force. An output which
peaks before the input sinusoid is described as leading.
A leading output might seem to violate cause and effect.
Shouldnt the effect follow the cause? There is no violation
of causality. The lead of an output does not occur on the first
cycle during the transient period. It develops during the transient period, and is established by steady-state, as will be
illustrated with the spring-mass-damper system.
The period of the sinusoid is calculated as T = 2 /w .
rad
Input Frequency w = 0.5
sec

T
T

0.93sec
= 0.074 decimal fraction of a cycle
12.6 sec

= 0.074 2 rad = 0.46 rad


= 0.074 360o = 26.5o
Input Frequency =1

T
T

0.78 sec
= 0.124 decimal fraction of a cycle
6.28 sec

= 0.124 2 rad = 0.78 rad


= 0.124 360o = 44.7o
Input Frequency =2

T
T

0.56 sec
= 0.18 decimal fraction of a cycle
3.14 sec
0.18 2 rad = = 1.13 rad
0.18 360o = = 64.8o

Phase angle between the input force and the output velocity increases with increasing input frequency. We shall see
that the theoretical limit for a first-order system is 90o of
phase shift . At very low frequency, the input force and the
resulting velocity of the mass move together with an imperceptible phase shift. As the input frequency increases, the
momentum and kinetic energy of the mass prevent it from
reversing direction as quickly as the input force, leading to a
phase shift between the input and output.
There is a slight error in the phase angle, at the input
frequency, =1rad/sec, due to limited numerical precision.
The input frequency equal to one over the time constant of
a first-order system, w = 1 / rad/sec, is known as the systems corner frequency, and equals 45, as we shall see,

when we discuss Bode plots in Sect.10.5. Determining the


center of the peak of noisy and misshapen sinusoids can
be more prone to error, than the numerical error found in
the evaluation of period and phase shifts from actual data.
Phase-angle data are very important in the process of determining the transfer function of a system from its frequency
response. Phase-angle data are useful, even though their
uncertainty is much greater than the corresponding magnitude data.
When input and output signals of a system driven by a sinusoidal input are viewed together on an oscilloscope, there
is nothing inherent in the image to distinguish a phase lag
from a phase lead. A shift in time between peaks of the two
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
must have more information, either the phase angles at other
input frequencies or the transient response. When frequency
response data are collected manually, begin with the lowest input frequency feasible. Increase the input frequency to
see the progression of the phase shift. An understanding of
theory is needed to interpret phase-angle data.

10.1.3Transient and Steady-State Response


of a Second-Order System
Figures10.2 and 10.3 show the sinusoidal response of second-order systems to be more varied than that of first-order
systems. The basis of the variation in sinusoidal response is
the internal energy transfers, or power flows, which occur
in a second-order system. The homogeneous, or natural,
response of a second-order system, apparent in its step response, is characterized by the systems eigenvalues. Two
real eigenvalues indicate an overdamped, non-oscillatory
homogeneous response. A complex conjugate pair of eigenvalues indicates an underdamped response. The same homogeneous, or natural, response of the second-order system is
superposed on its particular or steady-state response, which
is sinusoidal due to the sinusoidal input.
Overdamped second-order systems sinusoidal response
is analyzed as the superposition of the responses of two
first-order factors. The sinusoidal response of underdamped second-order systems is distinctly different. The
power flow between the two independent energy storage
elements, which produces oscillations in a step response,
leads to resonance in a sinusoidal response. Resonance
occurs when the input frequency approaches the natural
frequency of the system, due to enhanced power flow from
the sinusoidal source into the oscillating system. The maximum power flow into the system occurs, when the source
frequency matches the natural oscillation of the system.
The source pushes when the mass is naturally moving

10.1 Overview of Sinusoidal Excitation and Frequency Response

= 0.7

1.0

585

0.8

d
___
n 0.4

v1g

-1

0.2

0.4

0.6

Damping Ratio

0.8

away, and pulls when the mass is naturally approaching. A


person pushing a playground swing is synchronized with
the motion of the swing, pushing when the swing is moving away. Second-order systems with damping ratios less
than 0.7 exhibit a resonant peak, a local maximum in the
magnitude of their output variable, when the input frequency is nearly equal to their natural frequency.
We will first look at the envelope, which bounds the response of an underdamped second-order system, during
the transient period and into the steady-state period. Lets
use the parameters values, M = 2 kg, b = 2 N sec/m, and
K=200N/m. The spring-mass-dampers transfer function is
1
s
V1g ( s )
V1g ( s )
0.5s
M
=

= 2
b
K
F (s)
F ( s ) s + s + 100
s2 +
s+
M
M
The eigenvalues, s1 , s2 = j w , are s1 , s2 = 0.5 j 9.99.
The
ideal,
undamped,
natural
frequency
is
w n = K / M = 200 / 2 = 10. The damping ratio is
= 0.05. The resonant frequency of an underdamped system
is the actual, damped, frequency, d. The ideal, undamped,
natural frequency, n, is used for convenience. The difference between the damped frequency and the undamped natural frequency is less than 2% for the damping ratio <0.2,
and is one-tenth of 1% when the damping ratio =0.05,
Fig.10.13.
Figure10.14 shows the sinusoidal response of the springmass-damper system at the input frequencies of =8,
=10, and =12rad/sec. The bounding envelope for the
responses to the input frequencies of =8 and =12rad/sec,
Figs.10.14aandc, which are below and above the natural frequency, is a real exponential decay to a non-zero final value:
envelope(t ) = ( F0 G ( jw ) e t + F0 G ( jw ) )

-2

1.0

Fig. 10.13 Ratio of the observed, damped frequency, wd, over the ideal,
completely undamped, natural frequency wn plotted against damping

F0 |G(jw)|e + F0 |G(jw)|

m 0
___
sec

0.2
0

3
2

0.6

5 Steady-State

Transient

(10.5)

-3

t, sec

5 SteadyState

Transient

F0 |G(jw)|(1-e t )

v1g

=8

-F0 |G(jw)|e - F0 |G(jw)|

m 0
___
sec -2
-4
-6

= 10
0

t, sec

5 Steady-State

Transient

F0 |G(jw)|e + F0 |G(jw)|

v1g

m 0
___
sec -1
-2
-3

= 12

-F0 |G(jw)|e - F0 |G(jw)|


0

t, sec

Fig. 10.14 Velocity of the mass of the spring-mass-damper system


plotted with the envelopes, which bound the response. The parameter
values used were, b = 2 N sec/m, M = 2 kg and K=200 N/m. The
natural frequency, n, of the system is 10rad/sec. The frequency of
the input force was =8, =10, and =12 in a, b and c, respectively.
Note that the envelope in b is a growth to the steady-state value, while
the envelopes in a and c are decays to the steady-state values

where the final value is the magnitude of the steady-state


response. The coefficient of the real exponential is the real
component of the eigenvalues, .

10 Frequency Response

586

0.5

sin(10t)
0
+sin(11t)

v1g
m
___
sec

-1
-2

t, sec

Fig. 10.15 A beat results from the sum of two sinusoids of close to
equal frequencies. This beat is the sum of sin(10t) plus sin(11t)

The product in the bounding envelope equations, F0 G ( j w ) ,


is the steady-state magnitude of the output variable, the velocity of the mass, v1g. The steady-state amplitude of the
sinusoidal response of the system is due to the amplitude
of the input, F0, and its frequency, , and to the nature of
the system. The transfer function for the output variable,
G(s), is a mathematical description of the nature of the system. As it happens, the factor that the system contributes
to the magnitude of steady-state sinusoidal response is the
magnitude of its transfer function, evaluated as a conventional function by substituting for the Laplace variable, s,
the product of imaginary number j and the input frequency, , s=j. Evaluating G(j) yields a complex number,
which can be expressed in Cartesian form or in polar form
by its magnitude, G ( j w ) , and angle, G ( j w ) . The Frequency Response Equation is based on this mathematical
fact. It greatly reduces computational effort, when the only
aspect of the sinusoidal response of interest is the steadystate. We will review calculations with complex numbers
in Sect.10.2, and derive the Frequency Response Equation
in Sect.10.3.
The input frequencies, =8 and =12rad/sec, are close
enough to the systems natural frequency, n=10 rad/sec,
that a beat forms between 2 and 4sec into responses. A
beat results from the sum (superposition) of two sinusoids
of nearly equal frequencies. The resulting waveforms amplitude is bounded by a sinusoid of frequency equal to the
difference between the two sinusoids. The amplitude is
equal to the sum of the amplitudes of the two sinusoids,
Fig.10.15. We often hear beats, when rotating machinery is
running at approximately, but not precisely, the same speed.
Air travelers hear them, and may feel them, from the turbofan jet engines.

Steady-State

-F0 |G(jw)|e - F0 |G(jw)|


0

t, sec

= 20

8
5

Transient

At the resonant frequency, Fig.10.14b, the bounding envelope is a stable exponential growth:
envelopew n (t ) = F0 G ( jw ) (1 e t )
(10.6)

-F0 |G(jw)|(1-e

-0.5

Transient

m
___
sec

Steady-State

F0|G(jw)|(1-e t)

0.2

v1g

10

-0.2

-F0 |G(jw)|(1-e t )
0

= 30

t, sec

Transient

-F0 |G(jw)|(1-e t )

0.2

= 40

v1g

m 0
___
sec
-0.2

-F0 |G(jw)|e - F0 |G(jw)|


0

t, sec

Fig. 10.16 The transient portion of the spring-mass-damper systems sinusoidal response at input frequencies, w = 2w n , w = 3w n , and
w = 4w n , the first three harmonics

A second interesting phenomenon is the behavior of the


system, when the input frequency is an integer multiple of
its natural frequency. In the terminology of vibrations, the
natural frequency is the fundamental frequency, and its
integer multiples are its harmonics. The response of the
spring-mass-damper system to the first three harmonics of
its resonant frequency, =2n, =3n, and =4n, are
shown in Fig.10.16. The transient portion of the response
at =2n=20rad/sec is fit on the negative side by one

10.1 Overview of Sinusoidal Excitation and Frequency Response

bounding envelope, for both the resonant and non-resonant


cases. The positive side is almost flat. The responses at the
second and third harmonics are symmetric about zero.
As with the first-order, mass-damper system, the magnitude and phase shift in the steady-state response of the second-order spring-mass-damper system to a sinusoidal input
force can be understood by impulse acting to accelerate the
mass, the area under the trace of FM.
The period of the sinusoid is calculated as T = 2 /w .
rad
2
Input Frequency: w = 8
T=
= 0.785 sec
sec
w
The output, v1g, peaks before the input, F(t). The output
leads the input. The phase angle is positive.

T
T

+ 0.180 sec
= + 0.23 decimal fraction of a cycle
0.785 sec

587

a
2

v1g

m
___
sec

The input is the natural frequency of the system. The output, v1g, and the input, F(t) peak simultaneously. There is no
phase shift.
0
= No Phase Shift = 0
T
0.628 sec
rad
2
T=
= 0.524 sec
Input Frequency w = 12
sec
w
=

20

-2

=8
0

m
___
sec

t, sec

FM

F(t)

v1g

-35

100

T=0

50

2
0

-2

-50

-4
-6

= 10
0

t, sec

F(t)
N
FM
N

-100

5 Steady
-State

Transient

FM v1g

F(t)

40

20

v1g

0
-20

-1
o

The positive phase shift, or lead, for input frequencies less


than the natural frequency, w < w n , is due to differentiation
of the input force. Close inspection of Fig.10.17a reveals
that the output velocity of the mass does not lead the input
force on the first half cycle, preserving causality. The lead
of the output has started by the end of the first cycle. Differentiation contributes +/2 rad or +90 to the phase angle of
the output, over the entire frequency range. Without differentiation of the input, the output variable would lag the input,
i.e., have a negative phase shift, for all input frequencies.
The limit of the change of a second-order systems phase
angle is rad, or 180, as it progresses from low to high

-State

m 0
___
sec

0.112 sec
=
= 0.22 decimal fraction of a cycle
T
0.524 sec

FM

5 Steady
Transient

v1g

F(t)
N

-20

= 0.22 2 rad = 1.38 rad , = 0.22 360 = 79.2

35

The output, v1g, peaks after the input, F(t). The output
lags the input. The phase angle is negative.

v1g
F(t)

rad
2
T=
= 0.628 sec
Input Frequency: w = w n = 10
sec
w

-State

-1

FM

= + 0.23 2 rad = +1.45 rad , = + 0.23 360 = + 82.8o

5 Steady
Transient

-2

t, sec

FM
N

-40

= 12
0

F(t)
N

Fig. 10.17 Plots of the input force, F(t), the output velocity, v1g, and
the force acting to accelerate the mass, FM, of the spring-mass-damper
system. a Input force frequency w = 8rad/sec. The time shift is positive, t = + 0.18sec. The output leads the input. b Input force frequency
is the natural frequency, w = w n = 10 rad/sec. The time shift t=0 sec.
c Input force frequency w = 12 rad/sec. The time shift is negative,
t = 0.52sec. The output lags the input

frequencies. The change in the phase angle is /2 rad, or


90, at the second-order systems corner frequency, which
equals its natural frequency.

10 Frequency Response

588

A sin(t)

G(s)

A|G(j) |sin(t+)
where = G(j)

Fig. 10.18 Block diagram of the steady-state frequency response relationship

10.2 Frequency Response Relationship


The steady-state response of a linear system to a sinusoidal
input can be easily calculated from the systems transfer function, G(s). The relationship between the input and the output
sinusoids is expressed as a block diagram in Fig.10.18.
The effect of the system on the sinusoidal output is calculated, by evaluating the systems transfer function as a conventional complex function, substituting for s the product of
the imaginary number, j, and the input frequency, , in radians per second, where G ( jw ) is the magnitude of the transfer function, evaluated by substituting j for the Laplace variable, s, and G ( jw ) is the phase shift between the input
and output sinusoids. It cannot be overemphasized that the
frequency response relationship only holds in steady-state.
If the phase shift, , is negative, then the response follows,
or lags behind, the input. The system is classified as a lag
due to its phase lag. If is positive, the response precedes,
or leads, the input. The system is classified as a lead due
to its phase lead. Phase leads may seem to violate the principle of cause and effect, since the effect (output) appears to
occur before the cause (input). This is not the case, because
the frequency response equation only applies in steady-state.
The input has been applied long enough for the system to
get into a rhythm (or the swing of things, if you prefer).
The output is not being caused by the cycle that immediately
follows it, but by the cumulative effect of all the cycles that
preceded it, during the transient portion of the response.
The derivation of the frequency response equation is
presented in Sect.10.3. We will begin with an example calculation, by using the frequency response relationship. We
will then apply it to calculate the responses of the first and
second-order mechanical systems plotted above.

10.2.1Example Frequency Response


Calculation
The arithmetic of complex numbers was discussed in
Sect.2.7. It is reviewed briefly below. Calculations for the
Frequency Response Equation use both the Cartesian and
complex exponential forms of a complex number. Calculation of the angle (argument) and magnitude (modulus)
of a complex number presented in Cartesian form permits
the number to be expressed in complex exponential form.

10.2.1.1Example One: First-Order Mass-Damper


System
The transfer function is evaluated by substituting for the
Laplace variable, s, the product of the imaginary number,
j, and the input, or excitation frequency, . Any form of
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.
Poles are roots of the denominator, and zeros are roots of
the numerator. Note the constants, p and z, in the factors are
the opposites of the values of s, which represent the pole
and the zero.
1
V1g ( s )
s+z
M

=
Pole Zero Form G ( s ) = K p z
b
s+ p
F (s)
s+
M
An alternative form is the time-constant form
1

s + 1 V1g ( s )
b
Time Constant Form G ( s ) = K N

=
M
Ds +1
F (s)
s +1
b

G(s) is a complex function. The result of evaluating G(s)


for the purely imaginary argument jw , will be a complex number with the magnitude, G ( jw ) , and the phase
angle, G ( jw ).
Substituting in the parameters, M = 2 and b = 2, into the
pole-zero form yields:
G (s) =

V1g ( s )
F (s)

1
M

b
s+
M

G (s) =

1
2

2
s+
2

G (s) =

0.5
s +1

Evaluating the mass-damper systems transfer function,


G(s), for the product of the imaginary number, j, and the
input force frequency of w = 0.5 yields:
G ( j 0.5) =

0.5
j 0.5 + 1

Handheld calculators with engineering functions can evaluate this ratio, and return a complex number in Cartesian or
polar form, which is the efficient approach. We will express
the numerator and denominator as complex exponentials, to
illustrate the contribution of the numerator and denominator
to the magnitude and phase angle, Fig. 10.19.
Recall from Sect.2.7 the magnitude of a complex number
is calculated by the Pythagorean Theorem. The Pythagorean
theorem uses the magnitudes of the real and imaginary
components. Remove the imaginary unit vector, j, from the
imaginary component before squaring that component, to
avoid error.

10.2 Frequency Response Relationship

589

Im

Im
1+j0.5

j0.5

Re

Fig. 10.19 Plot of the numerator and denominator of the ratio


0.5
G ( j 0.5) =
j 0.5 + 1

An angle calculation is prone to error due to the signs of


the components of the complex numbers. The angle is calculated by using the inverse tangent function.
opposite
adjacent

= tan 1

Determine in which quadrant of the complex plane a complex


number lies, before calculating its angle. If a complex number
falls in the second or third quadrant, then the numbers angle
must be calculated indirectly, as illustrated in Sect.2.7. Handheld calculators inverse tangent function calculates the ratio
(opposite / adjacent ), and returns an angle in the first quadrant, if the ratio is positive, and in the fourth quadrant, if the
ratio is negative.
Express the numerator and denominator as complex exponentials. The numerators value is a positive real number.
Its magnitude is its value. Its angle with the positive real axis
is zero.
0.5 = 0.5 and 0.5 = 0 0.5 = 0.5e j 0
The denominators value is a complex number.

-50

-40

-30

-20

0.5
= 0.46 rad=26.4o
(1 + j 0.5) = tan 1
1
1 + j 0.5 = 1.03e

j 0.46

The complex number, G(j0.5), as a ratio of complex exponentials:


G ( j 0.5) =

0.5
0.5e j 0
=
j 0.5 + 1 1.03e j 0.46

Using the property of exponentials,

ea
= ea b :
eb

Re

-10

10
-j5

Fig. 10.20 Plot of the numerator and denominator of the ratio


j6
G2 ( j12) =
44 + j12

G ( j 0.5) =

0.5e j 0
0.5 ( j 0 j 0.46)
e
=
j 0.46
1.03
1.03e

G ( j 0.5) = 0.49e j 0.46

10.2.1.2Example Two: Second-Order SpringMass-Damper System


Evaluate the transfer function of the velocity of the mass
in the springmassdamper system for the product of the
imaginary number, j, and the input frequency, =12 rad/
sec. We will again use the parameter values M = 2 kg,
b = 2 N sec/m, and K=200 N/m.
1
s
M
=
G2 ( s ) =
b
K
F (s)
s2 +
s+
M
M
V1g ( s )

G2 ( s ) =

V1g ( s )
F (s)

0.5s
s 2 + s + 100

Evaluating for s = j12 :


G2 ( j12) =

0.5( j12)
j6
=
( j12) 2 + ( j12) + 100 144 + j12 + 100

1 + j 0.5 = 12 + 0.52 = 1.03


and

j10

j6

D
0.5

-44+j12

G2 ( j12) =

j6
44 + j12

Although we could divide the numerator and denominator by


two, we will not.
Express the numerator and denominator as complex exponentials, Fig. 10.20. The numerator is an imaginary number. Its magnitude is its value. Its angle with positive real
axis is /2 rad, or 90. Differentiation of the numerator adds
positive /2 rad, or 90, to the angle of a transfer function
evaluated for any input frequency, since the factor is on the
imaginary axis. Its magnitude changes with changing input
frequency, but its angle is fixed at 90.
j 6 = 6 and j 6 =

rad = 90o j 6 = 6e

10 Frequency Response

590

The denominator is a complex number. Its magnitude calculation is straightforward.


44 + j12 = 442 + 122 = 42.3
The complex number is in the second quadrant. A handheld
calculator would return the wrong inverse tangent for D:
12
tan 1
0.27 rad = 15.5o
44
This result is clearly incorrect. The desired angle is calculated, by finding the angle, , subtended by the complex
number, 44+j12 and the negative real axis. Subtract that
angle from , or 180.
12
= 0.27rad = 15.5o
44

= tan 1

D = = 0.27 = 2.87rad = 164.4o


( 44 + j12) D = 2.87rad = 164.4o

G (s) =

The denominator as a complex exponential is


44 + j12 = 42.3e j 2.87
The complex number, G2(j12), as a ratio of complex exponentials is
j

j j 2.87
6 e 2
G2 ( j12) =
= 0.14e 2
j 2.87
42.3 e
j

Transfer functions are unusual functions. So far, we


have used transfer functions as dynamic operators which
operate on the Laplace transform of the input function
multiplicatively, to yield the Laplace transform of the output function. Transfer functions can also be evaluated for
an argument, as are ordinary functions. A transfer function
is a complex function. Evaluating a transfer function for
a complex argument yields a complex number. In the frequency response equation, the systems transfer function is
evaluated by substituting for the Laplace variable s a purely
imaginary number, the input frequency multiplied by the
imaginary number j. The result is a complex number which,
like all complex numbers, can be expressed in terms of its
magnitude (or modulus) and angle (or phase angle or
simply phase).
The derivation of the frequency response equation requires the use of Laplace transforms, partial fraction expansion, and a representing complex functions as complex
exponentials. We will use a first-order system with the transfer function,

j j 2.87
6 e 2
2
0.14e
G2 ( j12) =
=
42.3 e j 2.87

G2 ( j12) = 0.14e j1.30 = 0.14e j 74.5

The phase angle of 74.5 is the correct phase angle for


this ideally linear, second-order system. The phase angle is
79.2, calculated from the traces of the input and output
variables in Fig.10.17c. Although the magnitude of the output is within 0.7% of its steady-state at a value of 5, the
phase angle is not.

10.3Derivation of the Frequency Response


Equation
A fortunate mathematical coincidence allows easy calculation of the steady-state response of a system to a sinusoidal
input. The relationship, known as the frequency response
equation, is a function of the frequency and amplitude of
the input sinusoid, and makes use of the systems transfer
function, Fig. 10.21.

1
Output ( s )
=
s+2
Input ( s )

and excite it with a sinusoid of amplitude A and angular frequency , input (t ) = A sin(w t ). The Laplace transform of the
input is
Aw
Input ( s ) = L { A sin(w t )} = 2
s + w2
We multiply the transfer function, G(s), by the Laplace
transform of the input, to obtain the Laplace transform of
the output.
Input ( s )G ( s ) =

Aw
1
= Output ( s )
s2 + w 2 s + 2

We are interested in the steady-state response of the system


to the sinusoidal input. We need to perform the inverse Laplace transform to return to the time-domain. We then take
the limit as t . The function, Output(s), is not in the
Laplace transform table, so we will need to perform partial
fraction expansion. Normally, we would keep the denominator factor, s 2 + w 2 , since sine and cosine are in Laplace transform tables. However, for this derivation, we will factor the
denominator into the product of first-order factors
Output ( s ) =
Output ( s ) =
Output ( s ) =

1
Aw
s2 + w 2 s + 2
Aw

( s + jw )( s jw ) s + 2
B1
B2
C
+
+
s + jw s jw s + 2

10.3 Derivation of the Frequency Response Equation

A sin(t)

G(s)

591

A|G(j) |sin(t+)

Using the same procedure, B2 and C are found to be

where = G(j)

B2 =

Fig. 10.21 Block diagram of the steady-state frequency response relationship

Aw
1
Aw
=
( jw + jw ) jw + 2 2 jw ( jw + 2)
B2 =

Determine the numerator constants by multiplying both sides


by the denominator of the corresponding term and canceling
identical factors. Evaluate by using the pole of that term. To
find the constant, B1,
Aw

( s + jw )( s jw ) s + 2

( s + jw ) =

C=

A
G ( jw )
2j
Aw
4 + w2

B1
B
C
( s + jw ) + 2 ( s + jw ) +
( s + jw )
s + jw
s jw
s+2

B2
Aw
1
C
= B1 +
( s + jw ) +
( s + jw )
w
w
2

+2
s
j
s
s
j
s
(
)
B2
Aw
1
C
= B1 +
+
( s + jw )
( s + jw )
s jw
s+2
( s jw ) s + 2 s = jw
s = jw
s = jw
B2
Aw
1
C
= B1 +
( jw + jw ) +
( jw + jw )
s+2
jw jw
( jw jw ) ( jw + 2)
B2
1
C
Aw
= B1 +
( jw + jw ) 0 +
( jw + jw )0
s+2
jw jw
( jw jw ) ( jw + 2)
Hence, the constant, B1, is
B1 =

where C is real. The expanded output function is

Aw
1
Aw
=
( jw jw ) ( jw + 2) 2 jw ( jw + 2)
B1 =

A
2 j ( jw + 2)

Output ( s ) =

which becomes more intimidating, when the complex constants are substituted in:

The transfer function is

Output ( s ) =
G ( s) =

1
s+2

Therefore, B1 can be written as


A
1
A
B1 =
=
G ( jw )
2 j ( jw + 2) 2 j

B1
B2
C
+
+
s + jw s jw s + 2

A
1
G ( jw )
s + jw
2j
A
1
Aw
1
+
G ( jw )
+
2
2j
s jw 4 + w s + 2

That B1 and B2 are complex constants does not affect the inverse Laplace transformation. Constants, either real or complex, can be factored out of the inverse transform, because it
is a linear operator.

L {Output (s)} = Output (t ) = L


1

B
Output (t ) = L 1 1 + L
s + jw

B1
B2
C
+
+

s + jw s jw s + 2
B2
1 C

+L

s + 2
s jw

1
1
1
1 1
Output (t ) = B1L 1
+ B2L
+ CL

s + 2
s + jw
s jw

10 Frequency Response

592

G ( jw ) is the systems transfer function, G ( s ), evaluated for


s = jw , where is the frequency of the forcing function.
The relationship between G ( jw ) and G ( jw ) can be seen,
when both are expressed as complex exponential:

All of the inverse transforms are of the form

1
at

=e
s
a
+

Therefore,

G ( jw ) =
Output (t ) = B1e

jw t

+ B2 e

j wt

+ Ce

2 t

Before dealing with the constants, B1, B2, and C, let us take
the limits of the exponentials as t . We will start with
the limit of e2t.
lim e 2t = e 2 =
t

G ( jw ) =

G ( jw ) =

Substituting in the complex constants, B1 and B2:


A
A
G ( jw )e jwt +
G ( jw )e jwt
2j
2j

We know that the steady-state response of the system, the


particular solution, will be a sinusoid, since we are forcing
the system with a sine. Our objective is to eliminate the complex terms, and yield a real trigonometric function. We will
need to use either Eulers sine or cosine equation.
e j e j
2j

or

cos( ) =

w2 + 4

w 2 + 4e

w
j tan 1
2

w
j tan 1
2

e j + e j
2

1 e j0

G ( jw ) =

jw + 2

w2 + 4 e

G ( jw ) =

Does e j w decay to zero? No. Recall that e j is the complex


exponential unit vector, where the angle of the unit vector is
the term in the exponent, multiplied by j. The complex exponential unit vector e jwt has a time-varying angle, = w t.
This unit vector rotates about the origin of its complex plane
completing a rotation of 2rad, as t increases by t = 2 / w .
The negative sign does not indicate decay; rather, it inverts
the sign of the angle of the unit vector. Consequently, the
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
angle rotating about the origin of a complex plane. However,
we cannot distinguish between and + 2 . Therefore,
there is no final value for either e jwt or e jwt .
Hence, as t , the steady-state output is

sin( ) =

G ( jw ) =

lim e j w t = e j w

Output (t ) steady state = B1e jwt + B2 e jwt + Ce 2t

1 e j0

and

1
1
= =0
e2

The homogeneous response of the system decays to zero in


steady-state. Regardless of the order of the system, we will
always have the same result. Only the particular, or forced,
response remains in steady-state.
Now, consider the limit

Output (t ) steady state =

1
G ( jw ) =
jw + 2

w2 + 4
1

w2 + 4

w
j tan 1
2

w
j tan 1
2

w
j tan 1
2

G ( jw ) and G ( jw ) have the same magnitude, differing only


in the sign of their angle. Let us generalize our notation by
defining
G ( jw )

w2 + 4

and

G ( jw ) = N ( jw ) D( jw )
w
w
= 0 tan 1 = tan 1
2

Using this notation with complex exponential unit vectors,


G ( jw ) = G ( jw ) e j

and

G ( jw ) = G ( jw ) e j

We can now rearrange the steady-state output expression


into a form which can be simplified, by using one of Eulers
equations
Output (t ) steady state =
Output (t ) steady state =

A
A
G ( jw )e jwt +
G ( jw )e jwt
2j
2j

A
A
G ( jw ) e j e jwt +
G ( jw ) e j e jwt
2j
2j

10.4 Fourier Series Approximation of Periodic Signals

Output (t ) steady state =

A
G ( jw ) e j e jwt e j e jwt
2j

593

e j (w t + ) e j (w t + )
Output (t ) steady state = A G ( jw )

2j

Output (t ) steady state = A G ( jw ) sin(w t + )

Input(t) 0
(10.7)

-1
0

Equation10.7 is the Frequency Response Equation which


we sought.

Although sinusoidal inputs are common and important to


consider in machine design, the frequency response equation
has even more general application. The response of a system
to a single sinusoid can be extended, to allow calculation of
the response of the system to any arbitrary periodic input.
The Fourier series allows any periodic signal f(t) to be constructed as the weighed sum (or superposition) of sinusoids,
plus a constant.

n =1

m =1

f (t ) = a 0 + an cos ( nw t ) + bm sin ( mw t )

(10.8)

where is the angular frequency (rad/sec). The cosines


and sines in the summation are integral multiples of the frequency, , of f(t), the periodic signal being replicated. The
frequency, , is called the fundamental frequency, even
though it is not necessarily the natural frequency of the system. The frequencies, which are integral multiples of , are
harmonics. If a dynamic system is linear, then the response
of the system to a periodic signal can be calculated, as the
weighted (or scaled) sum (or superposition) of the response
of the system to the sinusoids of the Fourier series of the
input. Although it takes an infinite number of sinusoids in
a Fourier series to exactly reproduce an arbitrary periodic
input, it takes relatively few to create a reasonably accurate
approximation.
The coefficients, a 0 , a n, and bm, are calculated by evaluating the following integrals

a0 =

an =

w
2

t +
1

f (t )dt

(10.9)

t1

t1

and


w
bm =

f (t ) cos(nw t )dt

10

t +
1

f (t ) sin(mw t )dt

(10.10)

(10.11)

t1

These integrals are intimidating. Recall how to calculate angular frequency from the period of a signal.

w=

2
T

This can be rearranged as

w 1
2
=
and
=T
w
2 T
The difference between the limits of integration is the period
of the fundamental frequency.
2
2

=T
t1 + t1 =
w
w
The integrals are evaluated over one period T, beginning at
any point in the cycle.
In the expression for a0, Eq.10.9, the integral of f (t )
with respect to time is multiplied by w / 2 = 1/ T . Hence,
the constant, a 0 , is the time average of the periodic signal,
f (t ). The constant a 0 is the vertical offset (also called a DC
offset) of f (t ) from zero.
The integrals for a n and bm , Eqs.10.10 and 10.11, yield
coefficients for the cosines and sines, inversely weighted for
the harmonic frequencies, n and m, respectively.
Recall from Calculus the integrals:
1

t +
1

t, sec

Fig. 10.22 Unit amplitude square wave with the period, T=2sec. Note
the tapered appearance due to the Heaviside unit step function being
assigned a value during its transition

10.4Fourier Series Approximation


of Periodic Signals

cos(nwt )dt = nw sin(nwt )

10 Frequency Response

594

0.5

Three
Term 0
Series

v1g
m
___
sec

F(t)

-1

-1
0

t, sec

10

Fig. 10.23 Three term Fourier series approximation of a unit amplitude square wave with a period T=2sec of Fig.10.22

t, sec

10

Fig. 10.25 Response of the first-order mass-damper system with the


parameter values, M=2 and b=2, to a unit square wave force input with
a period T of 2sec

0.5

Eight
Term 0
Series

v1g
m
___
sec

-1
0

t, sec

10

Fig. 10.24 Eight term Fourier series approximation of a unit amplitude


square wave with a period T=2 sec of Fig.10.22

and

-0.5

F(t)
0

N
-1

-0.5

t, sec

10

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 series approximation of a unit square wave force input with a period T
of 2sec

sin(mwt )dt = mw cos(mwt )


Consequently, the integrals for an and bm yield increasingly
small coefficients, as the integers, n and m, increase. The
Fourier series can be truncated, when the magnitude of the
coefficient is small enough, that the contribution of its term
to the sum is insignificant. In practice, relatively few terms
are needed in a Fourier series. Although it takes an infinite number of sinusoids in a Fourier series to exactly
reproduce an arbitrary periodic input, it takes relatively
few to create a reasonably accurate approximation. Consider a square wave with an amplitude of one, and a period
of T = 2 sec, which corresponds to a circular frequency of
w = rad/sec , Fig. 10.22.
Figures 10.23 and 10.24 are Fourier series approximations of this square wave, in which the series was truncated at
three terms and eight terms, respectively. Adding five higher
frequency terms to the series approximation will increase the
steepness of the transition, and flatten the waveforms top
and bottom. Note the corner of the approximate square wave.
A reasonable expectation would be for the oscillation at the
corners to vanish, when the number of terms in the series

is made reasonably large. That is not the case. The ringing at the corners of the approximate square wave is the
Gibbs phenomenon, named for the same J. Willard Gibbs
of Gibbs free energy, who identified the ringing as intrinsic
to a Fourier series approximation of a square wave using a
finite number of terms. Although the ringing at the corners
looks bad, in practice, it has little or no effect. The Fourier
series approximation of the square wave is accurate enough
for engineering applications.
Just how much fidelity do we need for a useful approximation? As it happens, in most applications, not much, because energy storage in systems is an integral process. This
is referred to as integral causality. Integration smoothens
sharp corners and obliterates fine details. Figure10.25 is the
response of the first-order mass-damper system with the parameter values, M=2 and b=2, to the unit square wave with
a period of T=2sec. The response of the same system to a
three-term Fourier series approximation of the square wave
is shown in Fig.10.26.
There is a slight wobble to the response to the three-term
Fourier series approximation. The transition is less sharp.
The differences between the response of the system to the

10.5 Bode Plots

595

0.4
0.2

v1g
m
___
sec

G(s)

A|G(j) |sin(t+)
where = G(j)

Fig. 10.28 Block diagram of the steady-state frequency response relationship

-0.2

A sin(t)

Square Wave
Response

Three Term
Series Response
0

0.5

t, sec

1.5

Fig. 10.27 Comparison of the first cycle of the response of the massdamper system to the unit square wave input force, dark trace, and the
three-term Fourier series approximation of the square wave force, light
trace

square wave and to the three-term Fourier series approximation are difficult to see, unless they are superimposed and
enlarged over the first cycle, Fig.10.27.
Although an approximate square wave constructed from
three sinusoids is certainly not an exact replica, the superposition of the systems response to the three sinusoids resemble the response of the system to the square wave, within
the precision of most engineering modeling.

10.5 Bode Plots


Except for resonance, the maximum values of the power
variables in an energetic system driven by a sinusoid input
occur during the transient responses of startup and shutdown. However, often our interest is the steady-state response of a system to a sinusoidal input, Fig. 10.28. When
we are only interested in the steady-state response, the plots
in the input and output sinusoids above have two shortcomings. First, it is difficult to extract the information we need
to predict the steady-state response of the system, G ( jw )
and G ( jw ), from these plots. Secondly, the information is
only for one specific input frequency. Bode plots, named
for Hendrik Bode (19051982), are plots of the magnitude
and phase angle of a system, over a wide range (many orders of magnitude) of input frequency. They are used to
represent experimental data, as well as design closed-loop
control systems and electronic filters to attenuate unwanted sinusoidal signals.
Bodes solution to the problem of presenting three dimensions or axes of data, G ( jw ) , G ( jw ), and , on the
two-dimensional plane of a piece of paper was to create
two two-dimensional plots which had one identical axis,
the frequency of the input sinusoid. The two Bode plots are
(1) a plot of the magnitude of the transfer function G ( jw )
on the ordinate (vertical axis), as either common (base 10)

log, or in decibels versus the common logarithm of the


input frequency on the abscissa (horizontal axis) and (2)
a plot of the phase angle G ( jw ) on the ordinate versus
common logarithm of the input frequency on the abscissa.
Note that the plot of the magnitude, G ( jw ) , is equivalent
to the response of the system to a sine with unit amplitude
since:
A G ( jw )
A

1 G ( jw )
1

= G ( jw )

10.5.1 Review of the Properties of Logarithms


Before we work with Bode plots, we will review the properties of logarithms. Common logarithms, base 10, and
natural logarithms, base e, Eulers number have the same
properties. The properties of logarithms derive from the
properties of the products and ratio of exponentials. Calculation of a logarithm and exponentiation are inverse operations.

( )

10log10 (a) = a and log10 10a = a


We are most familiar with exponents of base 10. When in
doubt regarding the property of an exponential or a logarithm, check it with an example in base 10.
The logarithm of a product is the sum of the logarithms.

10a 10b = 10a + b log 10a 10b = a + b


(10.12)
Similarly, the logarithm of a ratio is the difference of the
logarithms of the numerator and denominator.


10a
10a
= 10a b log b = a b
b
10
10

(10.13)

The logarithm of a power equals the power times the logarithm.




(10 )

a b

(( ) ) = a b

= 10a b log 10a

and

( )

log c d = d log ( c )

(10.14)

10 Frequency Response

596

The logarithm of a ratio is a restatement of the product of


logarithms.
10a
10a
= 10a 10 b log b = a b
b
10
10

Sound Intensity =

and

log c d 1 = log ( c ) log ( d )

(10.15)

Neither exponentiation nor the logarithm function is a linear


operator. Consequently, neither operation distributes onto a
sum.
10a + b 10a + 10b and log ( c + d ) log ( c ) + log ( d )
We noted that logarithms and exponentiation are inverse operations. The antilog of a quantity is the exponentiation of
that quantity to the base on which the logarithm was calculated. The antilog of a common logarithm is calculated by
raising 10 to that power.
antilog ( x ) = 10 x
An important logarithmic value is zero. The antilog of the
common logarithm equal to zero is one, 100 = 1. This result
derives from the properties of logarithms.
10
10
log = log (1) and log = log (10) log (10) = 0
10
10
log (1) = 0
Logarithms only exist for real numbers greater than zero.
There is no exponent for any base which yields either zero
or a negative number. Negative logarithms represent positive
real numbers between zero and one. They derive from the
logarithm of a ratio, where the numerator equals one:
a

Pressure of Sound Wave


Ambient Air Pressure

Sound intensity is presented in decibels, because the pressure ratio can vary over a very large range. It could be reported as the logarithm of the pressure ratio. However, to
avoid working with decimal fractions, the logarithm of the
pressure ratio, a bel, is multiplied by 10 to yield a decibel


y dB = log10 ( x )

(10.17)

Unfortunately, an exception was created for the logical definition of a decibel above. A common use of the unit decibel
is in electrical power calculations. Electrical power is the
product the current flowing through an element and the voltage across the element. Using v = iR. electrical power can
be expressed as:

P = iv P = i 2 R or P =

v2
R

Back in the bad old days of slide rules, any step which reduced the effort of calculations, even only slightly, was considered worthwhile. Consequently, since
v2
v
log i 2 R = 2 log (iR ) and log = 2 log
R
R

( )

it was decided to add a second definition for decibel when


only one of the two power variables is used! When the calculation involves only one power variable, rather than the
product of two power variables, a decibel is defined as
y dB = 20 log10 ( x )
(10.18)

10
10
1
= 10a b
= b = 100 b
b
b
10
10
10
and


familiar with decibel as the unit of sound intensity. In fact,


a decibel is not a unit. Sound is a pressure wave. Sound intensity is reported as a dimensionless ratio of air pressures.

1
log = log (1) log ( c ) = 0 log ( c ) = log ( c ) (10.16)
c

10.5.2Decibels
A decibel, abbreviated dB, is a scaled logarithm. The bel and
the decibel were first used in acoustics. The bel is no longer
used, but the decibel, 10 or 20 times a bel, is still widely used
in acoustics and electrical engineering. You are probably

The dual definitions can lead to miscommunication and


substantial error, particularly in this case, since the prefix,
deci, is Latin for 10. It is impossible to identify which of
the two definitions of decibel was used, when one is presented a value in dB without a context or the history of the
calculation. The definition of decibel used in system dynamics, control theory, and in Bode plots is y dB = 20 log10 ( x).
You may be wondering, Why use decibels at all? Why
not report the logarithm directly? That is a good question.
Scaling engineering values by factors of 10 and 100 is done to
eliminate the human errors in remembering and communicating decimal fractions. It is easier to remember a number between 1 and 100, than it is to remember a number between 0.1
and 0.001. We will use both decibels and straight logarithms.

10.5 Bode Plots

597

10

rad
, ___

100

1,000

10

100

1,000

sec

Fig. 10.29 Logarithmic interpolation value

Example Convert 3dB to a decimal fraction.

20

10

Fig. 10.30 Expansion of the interval from 10 to 20. Each interval is


scaled logarithmically

3
3
= log10 ( x ) 10 20 = 10log10 ( x)
20
0.15
10
= x = 0.708

3dB = 20 log10 ( x )

Verbally, an attenuation of 3 dB is a reduction of almost


30% (29.2%). This conversion example illustrates the reasoning behind scaling decimal fractions and the use of decibels. Which is easier to remember, 3dB or the base 10 log
of 0.15 ?

10.5.3 Interpolating on a Logarithmic Scale


Interpolating on a logarithmic scale was second nature to
engineers 40 years ago, prior to electronic calculators, when
calculations were performed with slide rules, where the
primary scales are logarithmic scales. (Multiplication and
division were performed by aligning the scales so as to either add or subtract logarithms. Slide-rule accuracy meant
three significant figures, where the third digit was estimated,
when one worked at the high end of the scale.) Fortunately,
slide rules are obsolete, but the use of logarithmic scales remains a practical way of presenting data which varies over
several orders of magnitude. Unless the datum one wishes to
read from a logarithmic plot falls on a grid line, reading data
from a logarithmic plot requires logarithmic interpolation.
The essence of interpolating on a logarithmic scale is the
division of a scale into logarithmic subdivisions. All logarithmic scales have the same proportions, regardless of the
magnitude of the number, which the scale may represent.
The technique is best presented as an example. First, we
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.
Now, the example interpolation. Determine the value of
the frequency, indicated by the dashed line on the logarithmic scale in Fig.10.29.
All intervals between gridlines on a logarithmic scale
are subdivided logarithmically. If gridlines were shown
in the interval from 10rad/sec to 20rad/sec, they would

LB
LA

10 20

100

1,000

Fig. 10.31 Measurement of position within the decade, 10100, to


determine the ratio, LA/LB

have the same relative spacing as the gridlines of a decade,


Fig.10.30.
Visualizing a decade grid in the interval from 10 to
20rad/sec would yield an estimate, hopefully accurate to
another digit, 16, in this case. If there were need for more accuracy or confidence, then measuring the lengths LA and LB,
Fig.10.31, using a linear scale, such as an engineers scale, if
you have one, or a millimeter scale, can be used to compute
the value as follows:
Interpolation = ( Lower Decade Limit ) 10

LA
LB

where the Lower Decade Limit refers to the decade in which


the interpolation is made, as in this example, the decade, 10
to 100. The ratio of the lengths LA over LB is the logarithm of
the unknown in that decade.
If the lengths are LA=7.0 and LB=32.5 on an arbitrary
linear scale, then the interpolation yields the value:
7
Interpolation = 10 10 32.5 = 10 (1.64) = 16.4

10.5.4 Log-Magnitude Bode Plots


Bode plots are created by evaluating the magnitude and
phase angle of a transfer function over a range of frequencies. The range chosen depends on the nature of the system.
As a rule of thumb, start with a reasonable frequency range,
or spectrum, by setting the lower and upper limits two

10 Frequency Response

598
Fig. 10.32 Log-magnitude Bode
plot of the velocity of the mass
of the first-order mass-damper
system

20

V1g(s)
0.5
_____
____
=
F(s)
s+1

log(0.5) = -0.30

log|G( j)| -1

-20

-2

-40

-3
0.01

orders of magnitude below and above either the frequency,


1
w , for a first-order system, or an overdamped second-

0.1

V1g ( s )
F (s)

1
M

b
s+
M

G1 ( s ) =

F (s)

0.5
s +1

An important reference on the log-magnitude axis is the logmagnitude=0 line. What is the value of any base raised to the
zero power, i.e., e0 or 100? One. The value of log G ( jw ) = 0
indicates that the magnitudes of the input and output variable
are equal.
log G ( jw ) = 0 antilog ( log G ( jw ) ) = 10

2 ( 1) 1
1
20 dB
=

or
rad 10 decade
decade

100

sec

rad
10

sec

A slope of 1 per decade increase in input frequency or,


equivalently, 20 dB per decade when the ordinate is in
decibels, for frequencies greater than the corner frequency,
is characteristic of the log-magnitude plot of a first-order
system.
Figure10.33 shows straight lines overlaid on the low frequency and high frequency portions of the trace, and extended until they intersect at the corner frequency, w = 1rad/sec.
The corner frequency of a first-order system is equal to the
magnitude of the systems pole

G1 ( s ) =

log G ( jw )

100 = 1 G ( jw ) = 1
Notice that between the lower input frequency limit of
w = 0.01 and w = 0.2,, the log-magnitude plot is horizontal or flat, with a value of log G ( jw ) = 0.30. The trace
curves downward over the frequency range of approximately w = 0.4 to w = 2. This is the knee of the trace.
We will identify the corner frequency in this range. The
plot appears to be virtually straight over the frequency range
above the knee to the upper frequency limit of the plot.
Notice that the trace passes through log G ( j10) = 1 and
log G ( j100) = 2. Again, the increase of frequency by a

-60
100

factor of 10 on a logarithmic scale is known as a decade.


Consequently, the slope of the trace in this frequency range is
slope =

V1g ( s )

10

rad
Input Frequency , ___
sec

order system and w w d or w n .

10.5.4.1First-Order System Log-Magnitude


Bode Plot
We will begin with the log-magnitude Bode plot of the firstorder, mass-damper system with the parameter values, M=2
and b=2, Fig. 10.32.

|G( j)|, dB

1
M
s+

b
M

wc =

b
M

(10.19)

10.5.4.2Second-Order System Log-Magnitude


Bode Plot
We next examine the log-magnitude plot of the second-order
spring-mass-damper system, Fig.10.34. We again use the parameter values, M = 2 kg, b = 2 N sec/m, and K=200 N/m.


1
s
V1g ( s )
0.5s
M
=
G2 ( s ) =
= 2
b
K
(
)
F (s)
F
s
+
s
s + 100
s2 +
s+
M
M
(10.20)

V1g ( s )

10.5 Bode Plots


Fig. 10.33 Log-magnitude Bode
plot of the velocity of the mass of
the first-order mass-damper system, annotated with straight lines
fit to the low and high-frequency
segments of the trace. The slope
of the high-frequency line is 1
or 20dB per decade of increase
in the input frequency. The intersection of the two straight lines
defines the corner frequency,
c=1 rad/sec

599

rad
___
Corner Frequency, c = 1 sec
1

V1g(s)
0.5
_____
____
=
F(s)
s+1

0
log(0.5) = -0.30

-1= -20 dB

log|G( j)| -1

-3
0.01

0.1

-20

V1g(s) _________
0.5s
_____
= 2
F(s)
s + s +100

-1

0
-20

-2

-40

-3

-60

-4

-80

-5
0.1

|G( j)|, dB

-60
100

10

rad
Input Frequency , ___
sec

log|G( j)|

-40

-2

Fig. 10.34 Log-magnitude Bode


plot of the velocity of the mass
of the second-order spring-massdamper system

20

10

100

1,000

|G( j)|, dB

-100
10,000

rad
Input Frequency , ___
sec
The slope of the low frequency line is +1, or +20dB, per decade of increase in the input frequency. The slope of the high
frequency line is 1, or 20dB, per decade of increase in the
input frequency. The intersection of the straight lines superposed on the low and high frequency portions of the trace
defines the corner frequency, c=10 rad/sec, Fig. 10.35.
To understand the Bode plot of this transfer function, we
will take the transfer function apart by using the properties of
logarithms, and then plot the individual factors.
a b
1
= log(a ) + log(b) + log
log
c
c
0.5s

log(G2 ( s )) = log 2
s + s + 100
1

log(G2 ( s )) = log(0.5) + log( s ) + log 2


s + s + 100

The log-magnitude Bode plots for the factors, 0.5 and s,


are shown in Fig.10.41. The factor, 0.5, is a constant not
a function of the input frequency. The factor, s, represents
differentiation of the input sine. The factor, log j , equals
zero, when = 1, because the magnitude of the imaginary
number, j, is unity. The slope of the trace is +1 or +20dB. An
increase in the input frequency by a factor of 10 is defined as
a decade. Since log(10) = 1, and log j is plotted against
frequency on a log scale, the slope is +1 per decade.
The constant factor, 0.5, is not a function of the input frequency, since it is a constant. Thus, its log-magnitude Bode
plot is a horizontal line with the value, log(0.5) = 0.30. Although multiplying a transfer function by 0.5 reduces, or attenuates, its magnitude, constant factors are termed gains
and given the symbol, K. A fractional gain is an attenuation, as in this case. The effect of a gain factor is a vertical shift of the log-magnitude Bode plot trace. If K>1, then

10 Frequency Response

600
Fig. 10.35 Log-magnitude Bode
plot of the velocity of the mass
of the second-order spring-massdamper system, annotated with
straight lines fit to the low and
high frequency segments of the
trace

rad
Corner Frequency, c = 10 ___
sec
0

V1g(s) _________
0.5s
_____
= 2
F(s)
s + s +100

-1

log|G( j)|

-2

0
-20
-40

-3

-60

-4

-80

-5
0.1

10

|G( j)|, dB

-1= -20 dB

+1= +20 dB

100

1,000

-100
10,000

rad
Input Frequency , ___
sec
Fig. 10.36 Log-magnitude Bode
plot of the transfer function factors, 0.5 and s. The logarithm of
the magnitude of j has a slope
of +1, or +20dB, per decade

rad
log|j|=0 at = 1 ___
sec

log|G( j)|

80

60

log|j|

+1= +20 dB

40

|G( j)|, dB

20

log|0.5|
0

log(0.5) = -0.30

-1
0.1

10

100

1,000

-20
10,000

rad
Input Frequency , ___
sec
log(K) is positive, and the vertical shift is upward by the
value of log(K). Conversely, if K<1, then log(K) is negative,
and the vertical shift is downward by the value of log(K).
The straight lines superposed on the low and high
frequency portions of the log-magnitude Bode plot,
1 / ( s 2 + s + 100), intersect at the corner frequency of
w c = 10 rad/sec. The corner frequency of an underdamped
second-order factor is its natural frequency. There is a distinct resonant peak in this log-magnitude plot, due to
the factors low damping ratio, =0.05. The low-frequency
value of the log-magnitude curve equals the log(1/100).
The straight line superposed on the trace for frequencies
greater than the corner frequency has the slope of 2 or
40dB per decade, twice the slope of the similar line of the
first-order transfer function, Fig.10.33.

Comparison of the slopes of the log-magnitude Bode plot


of the transfer function Eq.10.20, Fig.10.35, with the slopes
of the log-magnitude Bode plots of the three factors of that
transfer function, Figs.10.36 and 10.37 reveal that the slopes
of the factors sum to yield the slopes of the complete transfer
function. The summation of the slopes of the log-magnitude
plots of individual factors of a transfer function to yield the
slope of the log-magnitude plot of the entire transfer function
is due to the properties of logarithms,
a b
log
= log ( a ) + log (b ) log ( c ) log ( d )
c d
The effect of the damping ratio on the resonance of a secondorder system is illustrated by the family of log-magnitude

10.5 Bode Plots


Fig. 10.37 Log-magnitude Bode
plot of the transfer function factor 1/ ( s 2 + s + 100). The corner
frequency is w c = 10 rad/sec.
The straight line superposed on
the trace for frequencies greater
than the corner frequency has the
slope of 2=40dB per decade

601

rad
Corner Frequency, c = 10 ___
sec
0

1
___________
log
(j) 2 +j+100

( (

1
log ___
= -2
100

log|G( j)|

-40

-80

-4
-2= -40 dB

|G( j)|, dB

-120

-6

-8

0.1

10

100

1,000

-160

10,000

rad
Input Frequency , ___
sec
Fig. 10.38 A family of log-magnitude Bode plots showing the
effect of damping ratio on the
resonant peak. The input frequency is normalized (divided) by the
natural frequency of the system.
Note that the numerator of G(s)
is w n2 , giving the magnitude of
the transfer function the value
of one and the log-magnitude of
zero at low frequencies

1.5

30
= 0.02

1.0

20

= 0.05
= 0.1

10

0.5

= 0.2

log|G( j)|

dB

= 0.3
= 0.4

0.0

0.0

= 0.5
= 0.6

= 0.7
= 0.8
= 0.9
= 1.0

-0.5

-10

-20

-1.0

G(s) =
-1.5

0.1

n2
_____________
2
s + 2ns + n2
1

Normalized Input Frequency, __

-30
10

plots of Fig.10.38. These plots are for systems which have


the same ideal, undamped natural frequency n as the springmass-damper systems above but values of the damping ratio
which range from =0.02 to =1.0.
Notice that the amplification of the output for the damping ratio = 0.05 is log G ( jw d ) = 1.0, or a factor of 10.
Remember frequency response is the steady-state response

of a system to a sinusoidal input. Lightly damped systems


with resonant frequency below the frequencies excited by
their normal operating speeds can be operated safely, but
they must be started up deliberately, so as to pass through the
resonant frequency range quickly, relative to the duration of
the systems transient response.

10 Frequency Response

602
Fig. 10.39 Bode phase-angle
plot of the mass-damper system

rad
___
Corner Frequency, c = 1 sec
0o

V1g(s)
0.5
_____
____
=
F(s)
s+1

-15 o
-30

G(j) -45 o
-60 o
-75 o
-90 o
0.01

0.1

10

100

rad
Input Frequency , ___
sec
10.5.5 Phase-Angle Bode Plots
The phase angle of a transfer function can be calculated in
two ways. The first is to evaluate the transfer function in its
entirety. The second is to evaluate the numerator and denominator individually, then subtract the angle of the denominator from the phase angle of the numerator, to find the phase
angle of the transfer function.
G jw = N ( jw ) = N jw D jw
( )
( )
( )

D ( jw )

(10.21)

Bode phase-angle plots are one of the few instances in engineering, when dissimilar units for the same quantity are
used on the same plot. The units in this case are angular
measures. Quotations are placed around units, because all
angular measures are, in fact, dimensionless. A radian can be
seen as dimensionless from its definition: The angle which
subtends an arc of length, L, on a circle with radius R is

radians =

arc length L
radius R

arc length L length


[ radians] = radius R = length = dimensionless


It is not obvious that the angular measure of degrees is also
dimensionless, but it is. A degree is defined as the angle,
which subtends an arc length equal to 1/360 of the circumference of a circle. A degree is not an arc length. A degree is
a dimensionless ratio of lengths, an arc length over the circumference. Revolutions and cycles are also dimensionless
ratios, again, arc length over circumference.

In phase-angle Bode plots, it is common for the input frequency, , to be expressed in rad/sec, and the phase angle,
, to be expressed in degrees. The input frequency must be
in radians per second, to calculate either the magnitude or
the phase angle of the transfer function. The resulting phase
angle is then reported in the units we choose. Degrees have
two advantages over radians. It is much easier to visualize
an angle in degrees. Visualizing an angle in radians, other
than a common fraction or an integer multiple of , requires
a mental calculation to convert from radians to degrees or
a fraction of a circle or semicircle. Secondly, frequency response data are often used in graphical vector constructions.
Protractors are marked in degrees, not radians.

10.5.5.1First-Order System Phase-Angle


Bode Plot
We begin with phase-angle plot of the first-order massdamper system, still using the parameter values, M=2 and
b=2, Fig. 10.39.
We evaluate the frequency response of velocity of the
mass by making the substitution, s = jw .
V1g ( s )
F (s)

V1g ( jw )
0.5
0.5

=
s +1
F ( jw )
jw + 1

Recall from Sect.2.7.1.1, the difference between two vectors, Z1Z2, is the vector drawn from the tip of Z2 to the tip
of Z1, as shown in Fig.2.16, reproduced here for reference.
A complex number in Cartesian form is a sum. If we
express the sum, j+1, as a difference by a double-sign
inversion:
jw + 1 = jw ( 1)

10.5 Bode Plots

603

Fig. 2.16 Vector shortcut to construct the difference, Z3=Z1Z2.


The difference Z3 is drawn from
the tip of Z2 to the tip of Z1

Imaginary
-Z2 = -(3+j)

Z 1= -2+j2

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

-5

Z 3= Z 1 - Z 2

-4

-3

-2

s = j

_
j ___
+1

_
j ___
-(- _
1)
p = -1

-p = 1

Re

Fig. 10.40 Construction on a complex plane showing jw ( 1) and


jw + 1. The former allows the vector, s=j, to lie on the imaginary
axis, jw + 1 = jw ( 1)

then we can draw the complex number as a vector on the


s-plane with its tail on -1 and its tip on the imaginary axis,
Fig.10.40.
It is clear from Fig.10.40 that the maximum angle of the
first-order denominator factor, jw + 1, reaches a maximum of /2 = 90 as the frequency, , approaches infinity.
Because the factor is in the denominator, the angle contributed to the transfer function is negative.
V1g ( jw )

1
1
= lim
= lim
lim

jw + 1
jw ( 1)
F ( jw )

V1g ( jw )

lim
= lim 1 ( jw ( 1)) = = 90o

2
F ( jw )

3 Real

-j

repeated roots of the characteristic equation. The phase-angle plot of a critically damped system is Fig.10.41.

Z 2 = 3+j

-1

-Z2 = -(3+j)

Im

j2

All first-order system phase-angle plots have the same


shape. First-order systems transition through a change of 90
of phase angle, as the input frequency increases from approximately two decades below the corner frequency to two
decades above the corner frequency. They pass through 45
of phase-angle change at their corner frequency.
A damping ratio of =1.0 is the so-called critically
damped condition, with two identical real eigenvalues, the

10.5.5.2Underdamped Second-Order System


Phase-Angle Plot
The phase-angle plot of the second-order spring-massdamper 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,
the natural frequency, w c = w n = 10 rad/sec. The phase-angle plot begins at +90 at the lower input frequency limit.
Differentiation of the input, represented by the s in the numerator, adds +90 of phase angle, since s = jw is a purely
imaginary number, and the angle of the imaginary axis is
+90, measuring angles conventionally, counter clockwise
from the positive real axis.
The sharpness of the transition is due to the low damping ratio of =0.05. In the case of an almost completely undamped system, the transition is even sharper with almost
square corners. The response goes from having no phase
shift to being 180 out of phase with the sinusoidal input,
when the input frequency crosses the systems natural frequency. You may have experienced a system like this with
a lightly damped spring-mass-damper system at a childs
birthday party, a wooden paddle with the rubber ball tethered
to it by an elastic band. The ball and paddle move toward
each other, before the paddle strikes the ball, then away from
each other after the impact. The opposite directions of the
input, the velocity of the paddle, and the output, the velocity
of the ball, is 180 of phase shift.
Figure10.43 shows a family of phase-angle plots for underdamped second-order systems in the form of Eq.10.22,
with damping ratios varying from =0.02 to =1.0
G (s) =
G (s) =


( s w

w n2
s 2 + 2w n s + w n2
w n2

+ jw n 1 2

)( s w

jw n 1 2

(10.22)

10 Frequency Response

604
Fig. 10.41 Bode phase-angle
plot of a critically damped
second-order system with corner
frequencies, w c = 1rad/sec

rad
___
Corner Frequencies, c = 1 sec
0

-45

0.25
G(s) = __________

(s + 1)(s+1)

G(j) -90 o
-135

-180

0.01

0.1

10

100

rad
Input Frequency , ___
sec

Fig. 10.42 Bode phase-angle


plot of the spring-mass-damper
system

rad
Corner Frequency, c = 10 ___
sec
90

45

V1g(s) _________
0.5s
_____
= 2
F(s)
s + s +100

G(j) 0 o
-45

-90
0.1

10

100

1,000

10,000

rad
Input Frequency , ___
sec
The input frequency is normalized by the systems natural
frequency, which is also the corner frequency of an underdamped second-order system. The numerators angle is
zero, since it is a positive real number, and angles are measured from the positive real axis. The phase angles begin
at zero at the lower limit of the normalized frequency axis.
All of the phase-angle curves pass through a phase angle
of 90o, when the input frequency, , equals their natural
frequency, n.
The phase angle of an underdamped second-order factor can be understood by the constructions in a complex
plane, shown in Figs.10.44 and 10.45 of the denominator of Eq.1.21. Underdamped systems have complex

conjugate poles, which are symmetric relative to the real


axis. When the input frequency, =0, per Fig. 10.44a,
the angles made by the complex numbers, ( jw ( p1 ))
and ( jw ( p2 )), are equal, opposite, and sum to zero.
As the input frequency is increased, ( jw ( p1 ) ) becomes less negative. ( jw ( p1 ) ) equals zero, when the
input frequency equals the damped, natural frequency of
the system, w = w d . The magnitude, ( jw ( p1 )) , is at its
minimum at w = w d . If the damping ratio of the system
is less than =0.707, then the product of the denominator factors magnitudes, ( jw ( p1 )) ( jw ( p2 )) , is
also at its minimum when w = w d , and there is a resonant
peak in the magnitude plot. Figure10.46 shows rays from

10.6 Asymptotic Approximation of the Log-Magnitude Bode Plot


Fig. 10.43 A family of Bode
phase-angle plots of underdamped second-order systems
with damping ratios varying
from =0.02 to =1.0. The input
frequency, , is normalized by
the natural frequency, n

605
0o

-45 o

= 0.02
= 0.05
= 0.1
= 0.2
= 0.3
= 0.4

= 1.0
= 0.9
= 0.8
= 0.7
= 0.6
= 0.5

G(j) -90 o

-135 o

G(s) =
-180 o
0.01

0.1

b
-p1

jd= j n 1- 2

s-(-p1)

Im

2)

s = jd= jn 1-2

s-(

Re

Re

= - n

p
-(-

100

-p

(-p
s = j = 0
= - n

10

Normalized Input Frequency, __


n

Im

-p1

s-

Fig. 10.44 Constructions in a complex plane of


( s + p1 )( s + p2 ), written with
a double sign inversion as
( s ( p1 ))( s ( p2 )), and evaluated for s = jw . a s = jw = 0.
The angles are equal but opposite: 1 = 2 , b s = jw = jw d .
The angle, 1 = 0

n2
_____________
s 2 + 2s + n2

-p2

the origin representing loci of equal damping ratio. The


damping ratio, =0.707, is a ray inclined at 45 from the
negative real axis. For a given damped natural frequency,
the closer the pole is to the imaginary axis, the smaller
the magnitude, ( jw ( p1 )) , and the more quickly
( jw ( p1 ) ) changes, as the input frequency approaches and then exceeds the damped natural frequency.

10.6Asymptotic Approximation
of the Log-Magnitude Bode Plot
An understanding of frequency response is enhanced by the
ability to sketch the Bode log-magnitude and phase-angle
curves. Sketches of the log-magnitude plot will be reasonably

2
-jd = -jn 1-2

-jd = -jn 1-2

-p2

accurate, except near the corner frequencies of the first- and


second-order factors. Sketches of the phase-angle plot will
lack accuracy, but will be useful.
A transfer function is a complex function. When a transfer function is evaluated for s = jw , the result is a complex
number, or, equivalently, the ratio of two complex numbers,
Eq.1.22.


G ( jw ) =

(
(x

)( x
)( x

K xN1 + jy N1
D1

+ jyD1

G ( jw ) =

N2

D2

) (
) ( x

+ jy N2 xNm + jy Nm

+ jyD2

xN + jy N
= x + jy
xD + jyD

Dn

+ jyDn

)
(10.23)

10 Frequency Response

606

Rays of Equal
0.7

s = j > jd

0.4
0.2 0.05
0.6 0.5
0.3 0.1 j

s(-p

1)

Im

0.8

Moving a pole
toward the imaginary
axis decreases its 0.9
damping ratio

-p1

jd = jn 1-2
s-(-p

2)

0.95

Re

= - n

2
-p2

Fig. 10.46 Rays from the origin of the s-plane of equal damping ratio,

-jd = -jn 1-2

where N and D are the angles of the numerator and denominator. 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 .
As the input frequency approaches infinity, the angles of both complex 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 e j N
x + jy N j N jN
e e
G ( jw ) = N
j D
xD + jyD e
xD + jyD

G ( jw ) =

G ( jw ) =


In order to sketch Bode log-magnitude and phase-angle
plots, we will express the ratio of two complex numbers as
the product of the magnitude and complex exponential unit
vectors:


G ( jw ) =

s-plane

j x + jy

G ( jw ) =

(
(x

D1

+ jyD1

K xN1 + jy N1 e
xD1 + jyD1 e

j N1

j D1

j
e ( N D)

 xN + jy N
log
= log xN + jy N log ( xD + jyD
xD + jyD

(10.24)

(10.25)

(10.26)

We have the scheme we need. We will now apply it to a


transfer function with multiple factors in the numerator and
denominator.

)( x
)( x

K xN1 + jy N1

xD + jyD

The angles of the numerator and denominator are summed.


To create a sum of the magnitudes, we take the logarithm of
the ratio of the magnitudes.

x + jy N e ( N N )
xN + jy N e jN
xN + jy N
= N
=
xD + jyD
xD + jyD e jD
xD + jyD e j( xD + jyD )

G ( jw ) =

xN + jy N

N2

D2

) (
)( x

+ jy N2 xNm + jy Nm

+ jyD2

xN2 + jy N2 e
xD2 + jyD2 e

j N 2

j D2

Dn

+ jyDn

xNm + jy Nm e

j N m

xDn + jyDn e jDm

Rearranging to separate the magnitudes and the complex exponential factors:

G ( jw ) =

K xN1 + jy N1 xN2 + jy N2 xNm + jy Nm e


xD1 + jyD1 xD2 + jyD2 xDn + jyDn e

j N1

j D1

j N 2

j D2

j N m

e jDm

10.6 Asymptotic Approximation of the Log-Magnitude Bode Plot

607

We now use the fact that e a eb = e a + b :

G ( jw ) =

And the fact,

K xN1 + jy N1 xN2 + jy N2 xNm + jy Nm e


xD1 + jyD1 xD2 + jyD2 xDn + jyDn e

j D1 + D2 ++ Dm

overall gain of the transfer, and usually yields tremendous


errors. The asymptotic forms of first and underdamped sec-

ec
= ec d , which yields:
ed

G ( jw ) =

j N1 + N 2 ++ N m

K xN1 + jy N1 xN2 + jy N2 xNm + jy Nm


xD1 + jyD1 xD2 + jyD2 xDn + jyDn

The Bode log-magnitude and phase-angle plots separate


these factors as:

((

)(

j N1 + N 2 ++ N m D1 + D2 ++ Dm

ond-order factors have a resemblance which makes them


easier to remember; factor out a value equal to the constant

xN + jy N xN + jy N xN + jy N
1
1
2
2
m
m
log magnitude = log K

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

) (

Phase Angle = N1 + N2 +  + Nm D1 + D2 +  + Dm

10.6.1 Asymptotic Approximation Form


We will draw what are known as asymptotic approximations of the log-magnitude Bode plots. The straight lines,
which were superposed on the log-magnitude plots of the
first- and second-order systems, Figs.10.33, 10.35 and
10.37, are the asymptotes we will use to approximate the
log-magnitude curve. The term, asymptote, is used because
the low-frequency and high-frequency portions of the logmagnitude plots are not actually straight lines, although they
appear to be. Consequently, the approximation only equals
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
ends of the straight lines on the true curve.
The asymptotic approximation of the log-magnitude
Bode plot has required forms for both first and underdamped
second-order factors. First and underdamped second-order
factors must be put into the asymptotic approximation form
within the transfer function. Failure to do so will affect the

))

term, leaving unity behind.


Equation10.27 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
the denominator


G (s) =

(s + a)(s

Ks
2

+ 2w n s + w n2

(10.27)

The required forms of first- and second-order factors for the


asymptotic approximation have a common feature, i.e., the
constant term in the factor is unity. The asymptotic approximation form is created by factoring out the value of the constant term.
First-order factor asymptotic approximation form:


( s + a ) = a

s
+ 1
a

(10.28)

Underdamped second-order factor in asymptotic approximation form:


s 2 2w n s
(10.29)
s 2 + 2w n s + w n2 = w n2 2 +
+ 1
w n2
wn

10 Frequency Response

608

Expressing the transfer function G(s), Eq.10.27, in asymptotic form:


G (s) =
G (s) =

(s + a)(s

Ks
2

+ 2w n s + w n2

Ks
s 2 2w n s
s
a + 1 w n2 2 +
+ 1
a
w n2
wn

G (s) =

K
s
aw n2
2
2w n s
s s
+
1
+ 1

2 +
a
w n2
wn

The purpose of the asymptotic forms for first and secondorder factors becomes clear, when we consider the log-magnitudes of the factors in the limiting cases of low and high
frequencies.

10.6.2Asymptotic Approximation
of First-Order Factors
We will first consider the limiting cases of log-magnitude of

1
1
startfirst-order factor s
evaluated as log j w
+1

+ 1
a

a
ing with the low-frequency limit:
Low-frequency limit:

1
1
= log j 0
= log 1 = 0
lim log j w
w 0
1
+1
+1

The low-frequency limit establishes one of the two asymptotes. Recall the first-order transfer function for the massdamper system has the flat, or nearly horizontal lowfrequency portion of its log-magnitude curve, which corresponds to this asymptote. Recall the log-magnitude=0 line
means that at the input frequency, , the output variables
sinusoid has the same amplitude as the input sinusoid.
G ( jw ) =
log G ( jw ) = log

Output ( jw )
=1
Input ( jw )

Output ( jw )
= log (1) = 0
Input ( jw )

The general case of a first-order transfer function expressed


in asymptotic approximation form is:

GFirst ( s ) =
Order

K
s+ p

K
=
s+ p

K
1
=
s p s
+1
p + 1
p
K

K 1
log GFirst ( jw ) = log p jw
+1
Order
p
1
K
w
j
log GFirst ( jw ) = log
+ log
+1
p
Order
p
Consequently, the log-magnitude curve of a first-order transfer function has a constant gain factor, log( K /p ), which,
when summed to the first-order factors log-magnitude
curve, shifts the curve for the transfer function vertically.
The frequency at which the two asymptotes intersect is
called the corner frequency, even though it is not a right
angle. The corner frequencies of both first- and underdamped
second-order factors have physical significance. The corner
frequency of an underdamped second-order factor is the undamped natural frequency, n. What does the corner frequency
of a first-order factor correspond to, when a first-order factor
cannot oscillate? The time constant, , is the quantity, which
provides the units of time in the denominator of the corner frequency of a first-order factor. Recall that all angular measures
are dimensionless. Although we write the units of angular frequency, , as radians per second, the dimensions are
1

[w ] = sec =
c


K
K
K
1
G First ( s ) =

=
s+ p
s+ p p s
Order
p + 1
K 1
K
1
=

= wc = p
s + p p ( s + 1)

If the first-order factor is in the numerator of the transfer


function:
s
G First ( s ) = K ( s + z ) K ( s + z ) = z K + 1
z
Order
s
K ( s + z ) = z K + 1 w c = z
z
A common error is to confuse the constants, p and z, with
their corresponding pole and zero. The values of s which
yield zero in the denominator and the numerator are the pole
and zero, respectively:
G (s) =

s+z
s pole = p and szero = z
s+ p

10.6 Asymptotic Approximation of the Log-Magnitude Bode Plot

First-order log-magnitude plots all have the same shape. The


magnitude of the output is normalized, when it is divided
by the magnitude of the input. This enables us to obtain magnitude data for any input magnitude, and present the result,
as if the magnitude of the input were unity.
G ( jw ) =

Output ( jw )
Output ( jw )
=
Input ( jw )
Input ( jw )

The reverse process, that of scaling the normalized magnitude data to the magnitude of the input we expect, is one application of the log-magnitude Bode plot. Normalization and
the reverse process of scaling only work, if the system can be
reasonably modeled as linear. The further the system is from
an ideal linear system, the greater the error.
In the bad old days, templates with the knee curve of firstorder log-magnitude traces were sold at bookstores. Most
engineers calculated, or remembered, the attenuation of the
true first-order log-magnitude curve at the corner frequency,
3dB, and then blended in a smooth curve meeting the asymptotes, at frequencies of approximately 2.5 times above
and below the corner frequency.
The error between the log-magnitude approximation and
the true first-order log-magnitude trace at the corner frequency is the same in all cases, because the curve is normalized.
The approximation has no attenuation at the corner frequency, whereas the true value is
V1g ( jw c )
F ( jw c )

V1g ( jw c )
F ( jw c )
log

1
j +1

V1g ( jw c )
F ( jw c )

jw c + 1
=

1
j1 + 1

1
12 + 12

1
2

= 0.707

= log ( 0.707 ) = 0.151

V1g ( jw c )
F ( jw c )

We now evaluate the high frequency limit of log-magnitude of the first-order denominator factor:

1
1
1

lim log j w
log
=
= log
j

w
+
1
+
1
j

a
a
log

1
j

= log (1) log ( ) = 0 =

The high frequency limit of the log-magnitude is negative infinity. We need more information to establish the high-frequency
asymptote. We will determine the frequency at which the additive term, one, and the reciprocal factor, 1/a, become negligible,
relative to j. We evaluate the log-magnitude of the factor for
the input frequency equal to corner frequency, w = w c = a : v
1
1
1
log j w c
= log j a
= log
+1
+1
j +1
a
a
1
log
= log (1) log j + 1
j +1
log (1) log

( 1 + 1 ) = 0 log ( 2 ) = 0.1505
2

We now increase the frequency to 10 times the corner frequency, w = 10w c :


1
1
log j w
= log j 10w c
+1
+1
a
a
1
1
log j 10a
= log
= log (1) log j10 + 1
+1
j10 + 1
a
log (1) log

( 10 + 1 ) = 0 log (10.05) = 1.0022


2

and now to 100 times the corner frequency, w = 100w c :

In decibels,
20 log

609

= 20 log ( 0.707 ) = 20 ( 0.151) = 3.01dB

The attenuation of a first-order denominator factor at its


corner frequency, 3dB, provides the definition of bandwidth. The bandwidth of a signal is the frequency range,
over which the signal has less than 3dB attenuation, which
corresponds to an attenuation of 29.3%. The log-magnitude
plots of a first-order numerator factors and denominator factors are mirror images. Consequently, the error between the
asymptotic approximation, and the true log-magnitude curve
at the corner frequency is the inverse of the denominator factor. The true value of the magnitude at the corner frequency
of a numerator factor is an amplification of +3dB.

1
1

1
log j 100w c
= log j 100a
= log

+1
+1
j100 + 1
a
a
log (1) log j100 + 1 = log (1) log

( 100 + 1 )
2

1
= 0 log (100.005) = 2.00
log j 100w c
+1
a
The resulting log-magnitudes of the first-order factor,
evaluated at the frequencies of w = w c , w = 10w c , and
w = 100w c , are presented in Table10.1.
We see that the slope of the log-magnitude curve is
negative one per decade, between one and two decades in

10 Frequency Response

610
Table 10.1 Change in the log-magnitude of a first-order transfer function in the first two decades above the corner frequency
Change of a decade
Log(|G(j)|)
Input frequency
increase in frequency
c
0.1505
10c

1.0022

0.8517

100c

2.0000

0.9978

as that used for the first-order factor. We will evaluate the


low-frequency limit of the asymptotic approximation form,
determine the corner frequency, and then find the slope of the
high-frequency asymptote. Unfortunately, because the shape
of an underdamped second-order log-magnitude curve is a
function of the damping ratio, the asymptotic approximation
is significantly less accurate than for a first-order factor. In
the bad old days, one either used a template or attempted to
reproduce the resonant peak from a family of curves, similar
to Fig.10.38. In the present day, the ability to sketch the
log-magnitude curve is for understanding and design, not to
create accurate plots, which we now make using software.
The underdamped second-order asymptotic approximation form for a factor in the denominator of a transfer
function is

frequency above the corner frequency. We can now draw the


asymptotic approximation to the log-magnitude curve of a
first-order factor once we know the corner frequency.
If the first-order factor is in the denominator, the high-frequency asymptote slopes down at negative one per decade of
frequency from the corner frequency, Fig. 10.47. If the firstorder factor is in the numerator, the high-frequency asymptote slopes up at positive one per decade from the corner frequency. If the log-magnitude plot is scaled in decibels, then
the slopes are 20dB per decade and +20dB per decade.
Figure10.48 shows the error between the asymptotic approximation and the true log-magnitude curve for a first-order
denominator factor. A positive error means that the approximation is greater than the true magnitude. Note that the error
is symmetric about the corner frequency, c. Also note that the
error at the corner frequency equals the attenuation that defines the limit of a band width, log G ( jw c ) = 0.15 = 3 dB,
because the asymptotic approximation equals one, and its
logarithm equals zero at the corner frequency.

G (s) =

w n2
1
= 2
2
2
2w n s
s + 2w n s + w n
s
+
+1
w n2
w n2

Note the resemblance to the asymptotic approximation


form of a first-order factor. Both asymptotic approximation
forms have unity as the zero-order (constant) term of the
polynomial.
We calculate the low frequency limit of the magnitude:

lim

w 0

10.6.3Asymptotic Approximation of
Underdamped Second-Order Factors

( jw )
w

2
n

1
2

2w n jw

2
n

+1

( j 0)
w

2
n

2w n j 0

2
n

+1

=1

The low frequency limit is the same as for a first-order


factor. The low-frequency asymptote is the log-magnitude
equals zero line.

The development of the asymptotic approximation for an


underdamped second-order factor follows the same logic
Fig. 10.47 Asymptotic approximation of the log-magnitude of
first-order factors

1
_
Corner Frequency c =

log| j+1 |

1
log _____
j+1

60

40
+1

20

1 decade

1 decade

-1

-20

-1

-2
-3
0.1

dB

-40

10

100

Normalized Input Frequency, __


c

-60
1,000

10.6 Asymptotic Approximation of the Log-Magnitude Bode Plot


0.20

611

0.15

log

Error
0.10
log|G(j)|

0.05

( jw )
w

Error
dB

= log j w

0.1

0
100

10

We next consider the corner frequency for the asymptotic


approximation. Inspection of Fig.10.38, reveals that a line
representing the natural frequency of the system, /n=1,
passes through the resonant peak of the curves for damping
ratio, <0.1. We established by means of the s-plane vector
constructions, Figs.10.44, 10.45, and 10.46, that the resonant peak is a function of the actual, observed, damped frequency, d. We expect the asymptotic approximation to be in
error at the resonant peak of the system, in any case. Consequently, we will use the ideal, undamped, natural frequency,
n, as the corner frequency. It is a convenient value, since it
is the divisor used to create the asymptotic approximation
form of underdamped second order factors.
We now determine the slope of the high-frequency asymptote as we did for the first factor, by evaluating the logmagnitude at the corner frequencies, n, 10n and 100n.
Start with n:
1
+

2w n jw

w n2

+1

= log

( jw )
w

1
2

n
2
n

2w n jw n

w n2

wn

1
1
log
= log
j 2
1 + j 2 + 1
log

j 2

j wn

2
n

+1

= log (1) log (1) = 0 0 = 0

Success. Repeat the calculation for 10n:


log

( jw )
w

1
2

2w n jw

2
n

+1

w n2

= log

( j10w n )

1
2

log j 2 100w n2

w n2

1
210w n jw n

w n2

210w n jw n

w n2

w n2

+1

+1
1

= log j 100 w
2

w
log

log

1
2 w n j w n
log j w
= log 2
1
+
+
j + j 2 + 1
2
2

log

2
n

2
n

210 w n j w n

w n2

+1

1
2
= log (1) log 992 + ( 20 )

100 + j 20 + 1

+1

2
n

wn

( 2)(0.5) w n

2
n

and for 100n:

2
n

1
1
= log
j 2 + ( 2)( 0.5) j + 1
1 + j + 1

log

Fig. 10.48 Error between asymptotic approximation and true firstorder denominator factor log-magnitude curve. Positive error means
that the approximation was greater than the curve. The error curve is
inverted for a first-order numerator factor

w n2

w n2

( jw )

2
n

Normalized Input Frequency, __

log

( 2)(0.5) w n jw + 1

0
0.01

= log (1) log ( 2 ) = 0 log ( 2 ) = log ( 2 )

We can check our calculation by referring to Fig.10.38 for


the value of the damping ratio, , whose curve passes through
the log-magnitude equal zero line at the input frequency, n.
It is the curve with the damping ratio, =0.5.

( jw )
w

2
n

1
2

2w n jw

w n2

+1

= log

( j100w )

1
2

2
n

2100w n jw n

w n2

+1

1
= log j 10000 w
2

2
n

2
n

2100 w n j w n

w n2

1
= log
10000 + j 200 + 1
= log (1) log 9,999 + j 200 + 1
2
= 0 log 9,9992 + ( 200 )

+1

10 Frequency Response

612
Table 10.2 Change in the log-magnitude of an underdamped second-order transfer function in the first two decades above its corner
frequency
Input frequency Log(|G(j)|) Change of a decade increase in
frequency
n
0.3010
2.0043
1.7033
10n
4.0000
1.9957
100n

We must decide how to evaluate the terms with the damping


ratio, . One way to proceed is to evaluate the expressions
for =1, since that curve is a limiting condition farthest from
the high-frequency asymptote. Alternatively, we could use
=0.5, which lies closest to the asymptote. Lastly, we could
decide not to bother evaluating the terms with the damping
ratio, , because the negative real number under the radical
is larger in magnitude, and when squared, will dominate. We
will evaluate the log-magnitude expressions, using =1.
=n:

log

( jw )
w

2
n

1
2

2w n jw

2
n

+1

Fig. 10.49 Asymptotic approximation of an underdamped


second-order denominator factor,
superposed on the curves of the
true log-magnitude for damping
ratios from =1.0 to =0.02

= 0 log ( 2) = 0.3010

=10n:
log

( jw )
w

2
n

1
2

2w n jw

2
n

+1

= log (1) log

992 + 202

= 0 2.0043 = 2.0043
=100n:
log

( jw )
w n2

1
2

2w n jw

w n2

+1

2
= log (1) log 99992 + ( 200)

= 0 4.0000 = 4.0000
Summarizing the results in Table 10.2, we see that for =1,
the slope of the curve equals negative two per decade by
within a decade above the corner frequency. The equivalent
slope is 40dB per decade in decibels.
The asymptotic approximation of the log-magnitude curve
of an underdamped second-order denominator factor is plotted 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 =0.6. The

1.5

30
= 0.02

1.0

20

= 0.05
= 0.1

0.5

10

Asymptotic
Approximation

log|G( j)|

= 0.2

0.0

= 0.5
= 0.6

= 0.7
= 0.8
= 0.9
= 1.0

-0.5

-1.0

-1.5

dB

= 0.3
= 0.4

-10

-20

1
_____________
G(s) = 2
2s
s
__
+ ___ + 1
n2
n
0.1

Normalized Input Frequency, __

-30
10

10.6 Asymptotic Approximation of the Log-Magnitude Bode Plot


Fig. 10.50 Asymptotic approximation of an underdamped
second-order denominator factor.
The corner frequency is the undamped, natural frequency. The
slope after the corner frequency
is 2, or 40dB, for a denominator factor and +2, or +40 dB, for
a numerator factor

613

Corner Frequency c = n

( j) ____
2 j
log ___
+ +1
n
2n

1
____________
2
log (___
j) ____
2 j
+ +1
2
n
n

80

60

40

+2

20
1 decade

1 decade

dB

-20

-1
-2

-2

-40
-60

-3
-4
0.01

0.1

10

-80
100

Normalized Input Frequency, __


c

asymptotic approximations for underdamped second-order


factors in both the numerator and the denominator are shown
as Fig.10.50.

10.6.4 Integrator and Differentiator Factors


The Laplace variable, s, as a factor in the numerator, representing differentiation,
G (s) =

Ks

( s + a ) s 2 + 2w n s + w n2

and in the denominator, representing integration (a so-called


free integrator),
G (s) =

s ( s + a ) s 2 + 2w n s + w n2

if the input frequency is increased by a factor of 10, then


the log-magnitude of the differentiator increases by positive one, Fig. 10.51.
log j10w = log (10w ) = log (10) + log (w ) = 1 + log (w )
Hence, the slope of a differentiator factor is positive one, or
+20dB, per decade of frequency.
Integrators have the inverse relationship, since they are
factors of the denominator.
log

1
= log (1) log (w ) = 0 log (w ) = log (w )
jw

Increasing the input frequency by a factor of 10,


log

1
= log (1) log (10w ) = 0 log (10w )
j10w

contributes to the log-magnitude Bode plot across the frequency spectrum. Neither has a corner frequency.

log (10w ) = (log (10) + log (w )) = 1 log (w )

10.6.4.1Log-Magnitude of Integrator
and Differentiator Factors
A differentiator contributes the log of the input frequency to
the log-magnitude Bode plot

decreases the log-magnitude of the integrator factor by negative one. The slope of the integrators contribution to the
log-magnitude plot is negative one, or 20dB, per decade
increase in frequency.
Locating the straight log-magnitude line of a differentiator or integrator is easiest at either the lower limit of the input
frequency, or at the input frequency, w = 1. Note that the logmagnitude line of both the integrator and the differentiator

log jw = log (w )
Using the fact
log ( a b ) = log ( a ) + log (b ) .

10 Frequency Response

614

The Integrator and Differentiator cross the


rad
log| j| = 0 line at the input frequency = 1 ___
sec

Fig. 10.51 Log-magnitude


plots of an integrator and a
differentiator

log| j |

| |

1
log
j

60

40

20

-1

-20

-2

-40

-3
0.1

10

100

dB

-60
1,000

rad
Input Frequency , ___
sec
j
s-plane

Recall the property of exponentials,

s = j
s=

j = __
2

1
= e a
ea

1
1
1 j
=
= e 2
s jw w

(10.30)

Fig. 10.52 s-plane plots of a differentiator, s, and an integrator, 1/s

Factors in the denominator contribute negative angles to the


transfer functions phase angle. The angle of an integrator, a
multiplicative s in the denominator, is 90 = /2.
Although the magnitude of a differentiator and an integrator
is a function of the input frequency, their angles are not. The
angle contributed by a differentiator or an integrator is a constant 90o = /2 across the frequency range or spectrum.

cross the log-magnitude=0 line at the input frequency, =1,


since:

10.6.5 Gain Factors

1
__
=
s

3
__
1
__
=
j
2

1
__
=
s

1 - __
__
=
j
2

1 __
1
__
=
s
j

log j = log (1) = 0


and
log

1
1
= log = log (1) log (1) = 0 0 = 0
j
j

10.6.4.2Phase Angle of Integrator


and Differentiator Factors
The phase angle of a differentiator is easy to understand
from its plot on the s-plane, Fig.10.52. A purely imaginary
number, s = jw , lies on the imaginary axis at an angle of
+ 90o = + /2 to the positive real axis.
The phase angle of an integrator is understood, once it is
expressed as a complex exponential:
1
1
=
s jw

1
1
=
=
s
jw ejw

we

The term, gain, means multiplicative constant, which is


not necessarily positive or greater than one. The term, gain,
and the symbol, K, are used for many different constants. A
transfer functions gain is changed, when the transfer function
is expressed in asymptotic approximation form, as illustrated:
G (s) =
G (s) =

Ks

( s + a ) ( s 2 + 2w n s + w n2 )
Ks

s 2 2w n s
s
a + 1 w n2 2 +
+ 1
a
w n2
wn

K
s
G (s) = 2
2
w
a
n s s
2w n s
+ 1
+ 1 2 +
a
w n2
wn

10.6 Asymptotic Approximation of the Log-Magnitude Bode Plot


Fig. 10.53 Sketch One. Individual factors log-magnitude
plots

615

Corner Frequency c = 8
1

20

Gain = log(4)=0.6

0.6
0

12 dB
0

1
___
1s+1
__
8

1 decade

log|G(j)| -1

-1

Integrator

-20

dB

-1
1 decade

-2

-3
0.1

10

-40

100

-60
1,000

rad
Input Frequency , ___
sec

Here, the gain is the term,

K
.
aw n2

The log-magnitude of a gain is independent of frequency,


since a gain is a constant. Consequently, the log-magnitude
contribution of a gain constant is a horizontal line, because
the factor contributes equally across the frequency spectrum.
The effect of a gain on the resultant log-magnitude plot of
the entire transfer function is a vertical shift equal to the log
of the gain.

10.6.6Sketching Asymptotic Approximation


Log-Magnitude Bode Plots
Two sketches are need. One is for the log-magnitude plots
of the individual factors. The second is for the sum of individual factors of the first plot.
The process for sketching a log-magnitude Bode plot is
as follows:
1 Factor the numerator and denominator of the transfer function into: (1) gain K, (2) a differentiator or an integrator,
(3) first-order factors, and (4) underdamped second-order
factors.
2 Express the first and second-order factors of the transfer
function in the asymptotic approximation form.
3 Identify the corner frequencies of the first and underdamped second-order factors.
4 Estimate the frequency range by adding two decades
above the highest, and one decade below the lowest corner
frequency.
5 Sketch One. Sketch the individual factors contributions.

6 Calculate the log-magnitude range, by finding the difference between the sums of the log-magnitude of the factors at
the lower and upper frequency limits.
7 Sketch Two. Starting from the lower frequency limit, sum
the log-magnitudes, working from left to right.

10.6.6.1Example One: Sketch the Log-Magnitude


Bode Plot for
G1 ( s ) =

32
s ( s + 8)

1 The transfer function is already in factored form.


2 Express the factored transfer function in asymptotic approximation form.
G1 ( s ) =

32
=
s ( s + 8)

32
s
s 8 + 1
8

G1 ( s ) =

4
s

s + 1
8

3 The corner frequency of the first-order denominator factor


is 8rad/sec.
4 Estimate the input frequency range. The corner frequency
falls in the decade, 110rad/sec. Adding one decade below
and two decades above yields an input frequency range of
0.1 to 1,000rad/sec.
5 Sketch One, Fig.10.53. Sketch the individual factors contributions to the log-magnitude plot.
Calculate the log of the gain factor, Log(4)=0.60.
6 Calculate the range in log-magnitude of the sum of the
component factors, by summing the log-magnitude at the
lower and upper frequency limits.

10 Frequency Response

616
Fig. 10.54 Asymptotic approximation of the log-magnitude of the transfer function,
G ( s ) = 32/s ( s + 8)

Corner Frequency c = 8
2.0
1.6

40
32 dB

1 decade

-1
0

log|G(j)| -2.0

-40

1 decade

dB

-2
-4.0

-80
4
G1(s) = _________
s
s ___
+1
8

-6.0
0.1

10

-120
1,000

100

rad
Input Frequency , ___
sec
Log Magnitude at the Lower Frequency Limit:
1.0 + 0.6 + 0 = 1.6

10.6.6.2Example Two: Sketch the Log-Magnitude


Bode Plot for

Log Magnitude at the Upper Frequency Limit:

G2 ( s ) =

0.6 2.1 3.0 = 4.5


7 Sketch Two, Fig.10.54. Sum the factors. Start at the lower
frequency limit. Work to the right, in the direction of increasing frequency, to create the composite plot, which is the
asymptotic approximation of the log-magnitude plot of the
transfer function. The composite plot will consist of straightline segments. Sum the slopes of the lines of all factors, at
the lower frequency limit. The sum of the slopes is the slope
of the composite log-magnitude line, at the lower frequency
limit. The composite plot can be created by summing values
of the log-magnitudes of the individual factors at selected
frequency values. However, it is easier to sum slopes and
extend the log-magnitude line at that slope, until there is a
change in slope. The slope of the composite line from the
lower frequency limit does not change until the first corner
frequency. Therefore, extend the composite log-magnitude
line at the slope of 1/decade, from the lower frequency
limit to the corner of the first-order factor, w c = 8. The slope
changes at the corner frequency from negative one per decade to negative two per decade. There are no other corner
frequencies. Extend the composite log-magnitude line at the
slope of 2/decade from the corner frequency of the firstorder factor, w c = 8, to the upper frequency limit to complete the plot.
If we compare the asymptotic approximation and the
true log-magnitude curve, Fig. 10.55, we find the only
error is the frequency range of the first-order factors corner frequency.

32 ( s + 2)
2

s s + 30 s + 200

1 Although this transfer function appears to be factored,


check the eigenvalues of the second-order factor. If they are
purely real, then the factor is overdamped, and must be expressed as the product of the two first-order factors. Failure
to check the eigenvalues of second-order factors is the most
common error sketching log-magnitude Bode plots.
s 2 + 30 s + 200 = 0 s1 , s2 =

30 302 ( 4)( 200)

s1 = 10 and s2 = 20

The second-order factor is overdamped. It must be factored.


G2 ( s ) =

32 ( s + 2)
2

s s + 30 s + 200

32 ( s + 2)

s ( s + 10)( s + 20)

2 Express the factored transfer function in asymptotic approximation form.


s
+ 1
2
G2 ( s ) =
=
s ( s + 10)( s + 20)
s

s (10) + 1 ( 20) + 1
10
20
32 ( s + 2)

(32)( 2)

s
0.32 + 1
2
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


curve (black trace) and the
asymptotic approximation of
the log-magnitude (lighter
trace) of the transfer function,
G1 ( s ) = 32/s ( s + 8)

Corner Frequency c = 8
2.0
1.6

40
32 dB

Asymptotic
Approximation

log|G(j)| -2.0

-4.0

True Log
Magnitude Curve

-40

-80

4
G1(s) = _________
s
s ___
+1
8

-6.0
0.1

dB

10

-120
1,000

100

rad
Input Frequency , ___
sec
Fig. 10.56 Asymptotic
approximations of the logmagnitudes of the factors
of the transfer function,
G ( s ) = 32( s + 2)/( s ( s + 10)( s + 20))

Corner Frequencies c = 10
c = 20
c = 2
40

2.0

1.0

Integrator

__
1s + 1
2

1 decade

+1

-1

log|G( j)|

1
___
1 s+1
___
10

log(0.32) = -0.5

20

1 decade

-1

-1
1 decade

0.1

10

100

dB

-10

1 decade

Gain = 0.32

-1.0

-2.0

1
___

-20

1 s+1
___
20

1,000

-40
10,000

rad
Input Frequency , ___
sec
3 Identify the corner frequencies. The three first-order factors
have corner frequencies of w c1 = 2, w c 2 = 10, and w c3 = 20.
4 Estimate the frequency range. The lowest corner frequency
is in the decade, w = 1 to w = 10 . The highest corner frequency is in the decade, w = 10 to w = 100. The estimated
frequency range for the Bode log-magnitude plot is from
w = 0.1 to w = 10, 000.
5 Sketch One, Fig.10.56. Sketch the individual factors contributions to the log-magnitude plot.

Calculate the log of the gain factor. Log(0.32)=0.495=0.5.


6 Estimate the log-magnitude range of the summation. The
log-magnitude value at the lower frequency limit can be
summed directly. The upper limit summation requires extrapolating the lines of the individual factors, to estimate
their values for the input frequency, w = 10, 000 rad/sec. The
extrapolation is based on the slope of each line.
Log Magnitude at the Lower Frequency Limit:
1.0+00.5=0.5

10 Frequency Response

618
Fig. 10.57 Asymptotic
approximation of the logmagnitude Bode plot
of the transfer function,
G2 ( s ) = 32( s + 2)/( s ( s + 10)( s + 20))

Corner Frequencies c = 10
c = 20
c = 2
1.0

20
1 decade

-1
1 decade

-1.0

-20

-1

-40

-2.0

log|G( j)| -3.0

1 decade

-2

-4.0
-5.0

( )
( )( )

Fig. 10.58 True log-magnitude


curve (black trace) and the
asymptotic approximation of
the log-magnitude (lighter
trace) of the transfer function,
G ( s ) = 32( s + 2)/( s ( s + 10)( s + 20))

0.1

10

dB

-80
-100

s
___
+1
0.32
2
___________
-6.0 G2(s) =
s
s
___
s ___
+1
+1
20
10

-7.0

-60

-120
100

1,000

rad
Input Frequency , ___
sec

-140
10,000

Corner Frequencies c = 10
c = 20
c = 2
1.0

20
Asymptotic
Approximation

0
-1.0

-20

-2.0

-40
True Log
Magnitude Curve

log|G( j)| -3.0

-60

-5.0

-100

( )
( )( )

s
___
+1
0.32
2
___________
-6.0 G2(s) =
s
s
___
s ___
+1
+1
20
10

-7.0

3.70.52.7346.5
7 Sketch Two, Fig.10.57: Sum the factors to create the composite log-magnitude line for the transfer function. Work
from low frequency to high frequency. Sum the slope at each
inflection point.
Comparison of the asymptotic approximation and the true
log-magnitude curve, Fig.10.58, reveals greater error than

dB

-80

-4.0

Log Magnitude at the Upper Frequency Limit:

0.1

10

-120
100

1,000

rad
Input Frequency , ___
sec

-140
10,000

in Example One. The asymptotic approximation for a firstorder factor has minimum error, if there are no other corner
frequencies in the frequency band of the knee region of the
true curve. In this example, the two first-order denominator
factors knee regions overlap with each other and with knee
of the first-order numerator factor. This leads to an accumulation of error between the asymptotic approximation and the
true curve.

10.6 Asymptotic Approximation of the Log-Magnitude Bode Plot

619

Fig. 10.59 Asymptotic approximations of the log-magnitudes of


the factors of the transfer function,
G3 ( s ) = 132s /(( s + 2)( s 2 + 30 s + 400))

Corner Frequencies
c = 2 c = 20
1.0

20
Differentiator

+1
1
_________
s2 _____
30s
_____
+
+1
400 400

1 decade

log|G( j)|
log(0.165) = -0.78
-1.0

1
____
s
__
+1
2

-2.0

-3.0
0.1

10.6.6.3Example 3: Sketch the Log-Magnitude


Bode Plot for
G3 ( s ) =

( s + 2) ( s

132 s
2

+ 30 s + 400

1 Although this transfer function appears to be factored,


check the eigenvalues of the second-order factor.
s 2 + 30 s + 400 = 0 s1 , s2 =

30 302 ( 4)( 400)

s1 , s2 = 15.0 j13.2

The second-order factor has complex conjugate eigenvalues,


and is, therefore, underdamped. Keep the factor as a secondorder polynomial.
2 Express the transfer function in asymptotic approximation
form.
G3 ( s ) =
G3 ( s ) =

132 s

( s + 2) ( s 2 + 30s + 400)

132 s
s2

30
s
2 + 1 400
s + 1
+
2
400 400

G3 ( s ) =

0.165s
2

30
s s
s + 1
+
+ 1
2
400 400

1 decade

Gain = 0.165

dB
-15.6
-20

-2
-1
1 decade

10

100

-40

1,000

rad
Input Frequency , ___
sec

-60
10,000

3 Identify the corner frequencies. The first-order factor has


the corner frequency, w c1 = 2. The corner frequency of the
second-order factor is the ideal, undamped, natural frequency of the factor, w n = w c 2 = 400 = 20.
4 Estimate the frequency range. The lowest corner frequency
is in the decade, w = 1 to w = 10. The highest corner frequency is in the decade, w = 10 to w = 100. The estimated
frequency range for the Bode log-magnitude plot is from
w = 0.1 to w = 10, 000.
5 Sketch One, Fig.10.59. Sketch the individual factors contributions to the log-magnitude plot.
Calculate the log of the gain factor. Log(0.165)=0.783.
6 Estimate the log-magnitude range of the summation, extrapolating the lines to the upper frequency limit as needed.
Log Magnitude at the Lower Frequency Limit:
00.81.0=1.8
Log Magnitude at the Upper Frequency Limit:
40.83.75.45.9
The lower frequency sum estimate is too low an upper
limit for the log-magnitude axis. The log-magnitude line will
rise with a slope of positive one, due to the differentiator
from the lower frequency limit of w = 0.1 to the corner frequency of the first-order factor denominator factor at w = 2.
The upper limit must be at least log-magnitude equal zero.
There are eight horizontal divisions, so we will use log-magnitude equal zero as the upper limit.
7 Sketch Two, Fig.10.60 Sum the factors to create the composite log-magnitude line for the transfer function. Work
from low frequency to high frequency. Sum the slopes at
each inflection point.

10 Frequency Response

620
Fig. 10.60 Asymptotic approximation of the log-magnitudes of the transfer function,
G3 ( s ) = 132s /(( s + 2)( s 2 + 30 s + 400))

Corner Frequencies
c = 2 c = 20
20

1
0

1 decade

-20

-1 +1
-2

1 decade

log|G( j)| -3

-2

-40
-60

dB

-80

-4

-100
132s
___________
G3 (s) =
-6
-120
30s
s2
s + 1 _____
__
+ _____ + 1
400 400
2
-7
-140
0.1
1
10
100
1,000 10,000
-5

)(

rad
Input Frequency , ___
sec

Fig. 10.61 True log-magnitude


curve (black trace) and the
asymptotic approximation of
the log-magnitude (lighter
trace) of the transfer function,
G3 ( s ) = 132s / (( s + 2)( s 2 + 30s + 400)),
damping ratio = 0.75

Corner Frequencies
c = 2 c = 20
20

Asymptotic
Approximation

-20

-1
-2

-40

True Log Magnitude


Curve = 0.75

log|G( j)| -3

-60

-6

132s
G3 (s) = ___________
s2
30s
s + 1 _____
__
+ _____ + 1
400 400
2

-7
0.1

When comparing the asymptotic approximation with the true


log-magnitude curve, we find surprisingly good agreement,
per Fig.10.61. The surprising aspect is that the effect of
the damping ratio, , is not included in the asymptotic approximation. Since the systems damping ratio of =0.75 is
greater than =0.7, we may conclude the system log-magnitude plot has no resonant peak. If we reduce the damping
to =0.075, then we find substantial error in the frequency
range of the resonant peak, Fig.10.62.

dB

-80

-4
-5

)(

10

100

-100
-120

-140
1,000 10,000

rad
Input Frequency , ___
sec

10.6.7Bode Phase-Angle Plots of Transfer


Functions
We will now create phase-angle plots for transfer functions which have a combination of factors. The properties
of exponentials apply to complex exponentials, allowing the
phase-angle plots of each factor to be calculated separately.
Then the phase angles of the factors are summed, to yield the
phase-angle plot for the transfer function.

10.6 Asymptotic Approximation of the Log-Magnitude Bode Plot

621

Fig. 10.62 True log-magnitude


curve (black trace) and the
asymptotic approximation of
the log-magnitude (lighter
trace) of the transfer function,
G4 ( s ) = 132 s /(( s + 2)( s 2 + 3s + 400)),
damping ratio = 0.075

Corner Frequencies
c = 2 c = 20
1

20

True Log
Magnitude Curve
= 0.075

-1

-20
Asymptotic
Approximation

-2

-40

log|G( j)| -3

-60

-4

-80

-5
-6

-100

132s
G4 (s) = ___________
3s
s 2 _____
s + 1 _____
__
+
+1
400 400
2

)(

-7
0.1

10

dB

-120

-140
1,000 10,000

100

rad
Input Frequency , ___
sec

Corner Frequency c = 8

Fig. 10.63 Phase-angle Bode


plot of the factors of and the
complete transfer function,
G1 ( s ) = 32 / ( s ( s + 8))

45 o
0

-45

0.785

( )

1
__
s

o
G( j) -90

-0.785

Rad

-2.36
-3.14

-225
0.1

-1.57
G1(s)

-135o
-180

1
____
s+8

32
________
G1(s) =
s(s + 8)
1

10

100

-3.93
1,000

rad
Input Frequency , ___
sec
10.6.7.1 Example One: Phase-Angle Bode Plot
It is not necessary to express a transfer function in the asymptotic approximation form, in order to create its phase-angle Bode
plot, Fig. 10.63. It is important to factor the transfer function.
Each factor contributes to the angle of the transfer function.
G1 ( s ) =

32
G1 ( s ) =
s ( s + 8)

32
4
=
s
s
8s + 1 s + 1
8
8

s
G1 ( s ) = ( 4) s + 1 = 0 ( s ) + + 1

8
8

s
G1 ( s ) = 0 90o + + 1
8

10.6.7.2 Example Two: Phase-Angle Bode Plot


32( s + 2)
s ( s 2 + 30 s + 200)
Check the eigenvalues of the second-order factor. If the
eigenvalues are real, express the second-order polynomial as
the product of two first-order factors, Fig.10.64:

Phase-angle Bode plot of G2 ( s ) =

s 2 + 30 s + 200 = 0 s1 = 10 and s2 = 20
s 2 + 30 s + 200 = ( s + 10)( s + 20)
G2 ( s ) =

32 ( s + 2)
2

s s + 30 s + 200

32 ( s + 2)
s ( s + 10)( s + 20)

10 Frequency Response

622
Fig. 10.64 Phase-angle Bode
plot of the factors of and the
complete transfer function,
G2 ( s ) = 32( s + 2) / ( s ( s + 10)( s + 20))

c = 10
c = 20

Corner Frequencies c = 2

90

1.57

(s + 2)

1
( s_____
+ 10 )

1
( s_____
+ 20 )

( )
1
_
s

G( j)
-90

-180

( )
( )( )

0.1

Rad
-1.57

G2(s)

s
___
+1
0.32
2
___________
s
s
___
s ___
+1
+1
20
10

G2 (s) =

-3.14

10

100

1,000

rad
Input Frequency , ___
sec
Corner Frequencies c = 2

Fig. 10.65 Phase-angle Bode


plot of the factors of and the
complete transfer function,
G2 ( s ) = 132 s /(( s + 2)( s 2 + 30 s + 400))

90

c = 20
s

1.57
G3(s)

G( j) 0o

-90

1
_______
s 2 + 30s + 400

G 3(s) =

-180
0.01

1
____
s+2

Rad

-1.57

132 s
__________
(s + 2)(s 2 + 30s + 400)

0.1

10

100

-3.14
1,000

rad
Input Frequency , ___
sec

32 ( s + 2)
G2 ( s ) = 

s ( s + 10)( s + 20)

Check the eigenvalues of the second-order factor. The eigenvalues are complex conjugates. Keep the second-order
polynomial in its current form.

=  (32 ( s + 2))  ( s ( s + 10)( s + 20))

s 2 + 30 s + 400 = 0 s1 , s2 = j w d = 15 j13.2

= (  32 +  ( s + 2))

(  s +  ( s + 10) +  ( s + 20))

= ( 0 +  ( s + 2)) 90 +  ( s + 10) +  ( s + 20)

G2 ( s ) =  ( s + 2) 90o +  ( s + 10) +  ( s + 20)

10.6.7.3 Example Three: Phase-Angle Bode Plot


132 s
,
Phase-angle Bode plot of G3 ( s ) =
( s + 2)( s 2 + 30 s + 400)
Fig. 10.65.

Determine its ideal, undamped natural frequency, n, which


is its corner frequency, and damping ratio, :
s 2 + 30 s + 400 = s 2 + 2w n s + w n2

w n = 200 = 20 =

30
= 0.75
2w n

10.7 Nyquist Polar Plots

623

132 s
G3 ( s ) = 
2
( s + 2) s + 30 s + 400

180

=  (132 s )  ( s + 2) s 2 + 30 s + 400

= 132 0 +  s

G3 ( s ) = 90o  ( s + 2) +  s 2 + 30s + 400

90

))

 ( s + 2) +  s 2 + 30s + 400

))

arg(G( j)) 0o

))

-90

-180

0.1

The argument of a complex number is the primary, or


smallest angle needed to locate the direction of the complex number, when it is represented in magnitude and
angle, or magnitude and complex exponential unit vector
form. Unfortunately, the primary angle can be defined as
either 180o G ( jw ) 180o , or as 0o G ( jw ) < 360o.
Mathcads function arg() and MATLABs function angle()
both use the definition 180o G ( jw ) 180o. This is fine
for plotting the phase angle of individual factors, since the
maximum possible phase angle of a factor is 180. However, this definition creates errors, when the phase angle of a
transfer function exceeds 180. This situation occurs, when
the net order of the transfer function, i.e., the order of the
denominator minus the order of the numerator, is third order
or above. Erroneous phase-angle plots are easy to spot, because they have a step discontinuity of 180. For example,
the argument calculated by Mathcad of the fourth-order
transfer function

)(

s 2 + s + 4 s 2 + 20 s + 900

is plotted in Fig.10.66. To avoid errors, phase-angle plots


created in Mathcad or MATLAB must use superposition (or
summation) of the phase angles of the individual factors.
Fig. 10.67 Nyquist polar plot
of G ( s ) = 1/( s 2 + s + 3)

1,000

100

rad
Input Frequency , ___
sec

10.6.8 A Complex Numbers Argument

G (s) =

10

Fig. 10.66 Mathcads arg() function, and MATLABs angle() function


report the angle of a complex number as 180

10.7 Nyquist Polar Plots


Nyquist polar plots, named for Harry Nyquist (18891976),
present the frequency response of a system, by plotting the
result of evaluating transfer function, G(s), for s=j in Cartesian coordinates on the G(s) plane. Although we speak
of the complex plane, every complex function is entitled
its own plane to plot the result of evaluating it, hence the
G(s) plane.
Bode solved the problem of presenting three variables,
the input frequency, , G ( jw ) , and G ( jw ) , on twodimensional paper, by repeating the input frequency as the
abscissa on two plots. Nyquist used the alternative approach
of parameterizing (marking) the G(j) curve with the input
frequency. The axes of a Nyquist plot are the real and imaginary components of the transfer function being evaluated.
The transfer function is evaluated over a frequency range or
spectrum, as with Bode plots. The resulting Nyquist polar
plot curve can be interpreted as either Cartesian coordinates,
or as the trace of the tip of a complex exponential, represented by the vectors drawn from the origin of the G(s) plane
to the curve in Figs.10.67 and 10.68.
0.2

G(s) =

Im (G( j) ) 0

1
________
s2 + s + 3
=5

=0
= 0.5
=1

-0.2
-0.4
-0.6
-0.6

=2
= 1.5
-0.4

-0.2

Re(G( j))

0.2

0.4

0.6

10 Frequency Response

624
Fig. 10.68 Nyquist polar
plot of G ( s ) = s /( s 2 + s + 3)

0.6
=1

0.4
0.2

Im (G( j))

0
-0.2

= 1.5

= 0.5
=0

s
G(s)= ________
s2+ s + 3

= 10
=5

-0.4

=2

-0.6
-0.4 -0.2

Fig. 10.69 Portion of


the Nyquist polar plot of
G ( s ) = 1/( s 2 + s + 3), showing
the steady-state response to the
low and high input frequencies

0.4 0.6

=5

V1g ( s )
F (s)

s
s2 + s + 3

These values yield an underdamped system with the eigenvalues,


s1 , s2 = 0.5 j1.66 , the natural frequency, w n = 3 = 1.73,
and the damping ratio, = 0.289 .
An advantage of Nyquist plots is the magnitude and angle
of the transfer function evaluated for an input frequency are
shown to true linear scale. The true-scale magnitude is also
a disadvantage of Nyquist plots. The plot loses resolution,
when the magnitude becomes small. For example, upon
enlarging the Nyquist plot of G ( s ) = 1/ ( s 2 + s + 3), per
Fig. 10.69, all that can be determined above the input frequency of w = 10 , is that the curve approaches the origin at
an angle, 180o G ( jw ) 160o , as the input frequency
approaches infinity. Although we know that the angle of the
second-order transfer function approaches the origin at the
angle of 180, as the input frequency approaches infinity,

1.2

=0
= 0.5
o

G( j) = -180

s
G(s)= ________
2
s +s+3
0

and

1.0

= 10

-0.2

The Nyquist plots Figs.10.67 and 10.68 are of the transfer functions for the spring force and the mass velocity of
the spring-mass-damper system with the parameter values,
M=1, b=1, and K=3.

0.8

Re (G( j))

G( j0) = 0

Im (G( j))

FK ( s )
1
=
F (s) s2 + s + 3

0.2

0.2

Re (G( j))

=1

0.4

we could not estimate the angle of the final segment visually, without greatly enlarging the plot. In contrast, the logmagnitude Bode plot enhances small magnitude features,
which is very helpful, when using frequency response data
to determine a systems transfer function, i.e., system identification.
The small scale of Nyquist plots at high frequencies must
be weighed against the log scale of Bode plots producing a
false sense of precision. Although a frequency of 10rad/sec
is usually well within the capability of experimental apparatus, frequency response data are limited by the power of the
input (or excitation), and the signal to noise ratio of the output variable. Depending on its mass, or mass moment of inertia, an input frequency of 10rad/sec may be at the limit of
experimental feasibility for a mechanical system described
by the transfer function, G ( s ) = 1 / ( s 2 + s + 3).
A second drawback of Nyquist polar plot is that plots of
numerator and denominator factors not only lack the symmetry found in Bode plots, they do not even resemble one
another. A plot of a factor, say, (s+1), is completely different, if the factor is in the numerator rather that the denominator. It is only feasible to sketch Nyquist plots for very
simple cases. Recall that Bode plots contain mirror image

Summary

625

Fig. 10.70 Nyquist polar plots.


a Differentiator, G ( s ) = s.
b Integrator, G ( s ) = 1/s

Im (G( j))

-1

=4
=3
=2
=1
=0
-2

Re (G( j))

-4

= 0.2
-6
-8

= 0.1

-10
-2

Re (G( j))

0.4

=4

Approaches the origin at -90


as approaches infinity

0.2

G(s) = s + 1

Im (G( j)) 0

=2

Re (G( j))

symmetry of numerator and denominator factors for both


the log-magnitude and phase-angle plots. The contribution,
a factor in the numerator, is the mirror image, relative to the
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
Nyquist plots.
The only numerator and denominator factors with similar
shapes are differentiators and integrators, Fig.10.70, since
they both appear as straight lines on the imaginary axis. They
do not have mirror image symmetry. They are reciprocals.
The magnitude of the integrator approaches infinity, as the
input frequency approaches zero.
The Nyquist plots for first-order factors, Fig.10.71, and
second-order factors, Figs.10.72 and 10.73, share similarities, in that the magnitudes of the numerator factors head
to infinity with increasing frequency, as the magnitudes of
the denominators head to zero. The first-order denominator

-0.8
-1.0
-0.2

= 0.1
= 0.25
= 0.5

=1

-0.6

= 0.5
= 0.25 = 0
0

=4
=2

-0.4

=1

=0
=10

-0.2

Im (G( j))

=5

Im (G( j))

= 1 = 0.75
= 0.5
-2
= 0.4
= 0.3

= 7.5

Fig. 10.71 Nyquist polar plots


of first-order factors. a Numerator factor, G ( s ) = s + 1. b Denominator factor, G ( s ) = 1/ ( s + 1)

= 10

10

1
G(s) = __
s

G(s) = s

G(s) =
0

0.2

1
____
s+1

0.4 0.6 0.8 1.0

Re (G( j))

1.2

factor approaches zero at 90, and the second-order factor


approaches zero at 180, respectively, as factors are dominated by the frequency term.

Summary
Sinusoidal inputs to energetic systems are common. Wave
phenomena and rotation provide the ultimate source of sinusoids. Any periodic input can be represented as a summation
of sinusoids at different frequencies with the Fourier series.
Both the transient and steady-state response of an energetic system to a sinusoidal input are important to consider
in machine design.
Transient Sinusoidal ResponseThe maximum or minimum value of the power variables in a system not subject
to resonance occurs during the transient period of sinusoidal

10 Frequency Response

626
Fig. 10.72 Nyquist polar plot of
the second-order numerator factor, G ( s ) = s 2 + 4s + 64

80

G(s) = s2 + 4s + 64

60
= 10

Im (G( j)) 40

= 7.5
=5
=4

20
0
-80

0.005
=0
=1
=2
=3

Im (G( j))

Approaches the origin


=4

at -180 when

-0.01

=5

= 10

1
__________
G(s) = 2
s + 4s + 64

-0.02
Crosses imaginary axis
at -90 when = n

-0.03
= 7.5

-0.01

0.01

Re (G( j))

0.02

Fig. 10.73 Nyquist polar plot of the second-order numerator factor,


G ( s ) = 1/( s 2 + 4 s + 64)

A sin(t)

G(s)

A|G(j) |sin(t+)
where = G(j)

=2
=3
=1 =0

-60

-40

-20

Re (G( j))

20

40

60

80

Bode plots and Nyquist polar plots are plots of the steadystate frequency response of a system. The corner frequencies
of first-order factors are the inverse of their time constant,
w c = 1/ , which equals the magnitude of the pole, or zero,
of the factor. The corner frequency of a second-order factor is its natural frequency. The phase angles of integrators
and differentiators are constants, 90 and +90, respectively. The maximum change, or transition, of a first-order
factors phase angle is 90 and of a second-order factor is
180. Both the first- and second-order phase-angle curves
pass through half of their total transition at the factors corner frequency. The phase angles of first-order factors have a
single shape, allowing them to be sketched accurately. The
shape of the phase-angle Bode plot of a second-order factor
is a function of the damping ratio. It requires an exemplar to
sketch accurately. Bode log-magnitude plots can be sketched
to reasonable accuracy by using asymptotic approximation,
except that resonant peaks, if any, are omitted.
Nyquist plots are true, linear sketches of the systems
transfer function, evaluated for s=j. They must be parameterized by the input frequency. They cannot be sketched accurately.

Fig. 10.18 Steady-state frequency response relationship

excitation. Many failures occur during startup and shutdown


of systems. Calculation of the transient response of a system to a sinusoidal input is usually accomplished with either
Laplace transforms or state-space.
Steady-State ResponseThe steady-state response of a
system to a sinusoidal input is easy to calculate from the
systems transfer function, using the frequency response
relationship, Fig.10.18.
The frequency response relationship only applies to the
steady-state response. It represents only the particular response of a system to a sinusoidal input. The natural or homogeneous response is not included.

Problems
Problem 10.1 Fig P10.1 shows a translational mechanical
system. Mass M is supported by frictionless rollers. A dashpot with damping b and a spring with spring constant K are
attached between the mass and ground. The mass is acted
upon by a sinusoidal force, F(t). The parameter values are
M=3, b=2, and K=100
10.1.a Derive the system equation for:
i The velocity of the mass.
ii The force acting through the spring.
iii The force acting to accelerate the mass.
10.1.b Derive the transfer function for the system equation
of part a and plot the poles and zeros (if any) of transfer function on the s-plane.

Problems

627

x,v

F(t)

b
v(t)

M
K
Frictionless
rollers

Fig.P10.1 Translational spring-mass-damper system acted on by the


force, F(t)

Shaft rigidly attached


Torsion spring K
Flywheel
rotational inertia J

Torque
input T(t)
Hydrodynamic
bearing damping b
Fig. P10.2 Rotational system consisting of a torque source, a flywheel
with damping, and a torsion spring

10.1.c Determine the steady-state response of the transfer


functions of part b for:
i F (t ) = 10 N sin(3t )
ii F (t ) = 10 N sin(6t )
iii F (t ) = 10 N sin(12t )
10.1.d Formulate the state-space representation of the system with the output variables:
i The velocity of the mass.
ii The force acting through the spring.
iii The force acting to accelerate the mass.
10.1.e Solve the state and output equations of part d for the
sinusoidal inputs of part c for the duration, t 10 ,
1
.
where =

min

Problem 10.2 Fig. P 10.2 shows a rotational mechanical


system. The flywheel has rotational inertia J. The flywheels
shaft is supported by a hydrodynamic bearing with damping b and attached to a torsion spring with spring constant
K. The other end of the torsion spring is rigidly attached to
the machine frame, which is ground. The system is driven a

Fig. P10.3 RLC electric circuit

sinusoidal torque, T(t). The parameter values are J=5, b=4,


and K=1,000.
10.2.a Derive the system equation for:
i The angular velocity of the flywheel.
ii The torque acting through the spring.
iii The torque acting to accelerate the flywheel.
10.2.b Derive the transfer function for the system equation
part a and plot the poles and zeros (if any) of transfer
function on the s-plane.
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 )
10.2.d Formulate the state-space representation of the system with the output variables:
i The angular velocity of the flywheel.
ii The torque acting through the spring.
iii The torque acting to accelerate the flywheel.
10.2.e Solve the state and output equations of part d for the
sinusoidal inputs of part c for the duration, t 10 ,
1
.
where =

min

Problem 10.3 Fig. P. 10.3 shows an RLC electric circuit


with a voltage source. The input is a sinusoidal voltage, v(t).
The parameter values are L = 0.01 H, C = 4 10 6 F, and
R = 0.1 .
10.3.a Derive the system equation for:
i The current through the inductor.
ii The voltage across the capacitor.
iii The current from the voltage source.
10.3.b Derive the transfer function for the system equation
of part a and plot the poles and zeros (if any) of transfer function on the s-plane.
10.3.c Calculate the ideal, undamped, natural frequency, n,
and the damping ratio, , of the circuit.
10.3.d Determine the steady-state response of the transfer
functions of part b using the natural frequency determined in part c for:
wn

i v(t ) = 24 VDC sin t
2
ii v(t ) = 24 VDC sin(w n t )
iii v(t ) = 24 VDC sin(2w n t )

10 Frequency Response

628

C
P(t)
vent to
atmosphere

R1

Pump

I
R2

Fluid
Reservoir

Fig. P10.4 Schematic of a fluid system:

10.3.e Formulate the state-space representation of the system with the output variables:
i The current through the inductor.
ii The voltage across the capacitor.
iii The current from the voltage source.
10.3.f Solve the state and output equations of part e for the
sinusoidal inputs of part c for the duration, t 10 ,
1
.
where =

min

Problem 10.4 Fig. P10.4 shows a fluid system with a pump


modeled as a pressure source, two fluid resistances, a fluid
inertance, and a fluid capacitance (accumulator). The parameter values are:
R1 = 1

MPa sec 2
m3
MPa
MPa
5
I
=
4
C
=
20
R
=
2
m3
MPa
m3
m3

10.4.a Derive the system equation for


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.
10.4.b Derive the transfer function for the system equation
of part a and plot the poles and zeros (if any) of transfer function on the s-plane.
10.4.c Calculate the ideal, undamped, natural frequency, n,
and the damping ratio, , of the fluid system.

Fig. P10.5 Schematic of a DC


motor with an inertial load and
driven by a voltage source

10.4.d Determine the steady-state response of the transfer


functions of part b using the natural frequency determined in part c for
wn
i p (t ) = 3, 000 psi sin t
2
ii p (t ) = 3, 000 psi sin(w n t )
iii p (t ) = 3, 000 psi sin(2w n t )
10.4.e Formulate the state-space representation of the system with the output variables:
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.
10.4.f Solve the state and output equations of part e for the
sinusoidal inputs of part c for the duration, t 10 ,
1
.
where =

min

Problem 10.5 An electromechanical schematic of a DC


motor is shown in Fig. P10.5. The motors resistance is
R = 4 . The relationship between the motor current and the
motor torque is TM = KT iM , where KT = 8 N m/A. The motors mass moment of inertia is J M = 0.3 kg m 2 . The motor
turns a flywheel with mass moment of inertia, J L = 2 kg m 2,
and damping b = 0.1 N m sec/ rad .
10.5.a Derive the system equations for:
i. The current from the source.
ii. The back EMF.
iii. The angular velocity of mass moment of
inertia,JL.
and check their units.
10.5.b Derive the transfer function for the system equation
of part a and plot the poles and zeros (if any) of transfer function on the s-plane.
10.5.c Determine the steady-state response of the transfer
functions of part b for:
i v(t ) = 24 VDC sin(3.5t )
ii v(t ) = 24 VDC sin(7t )
iii v(t ) = 24 VDC sin(14t )
10.5.d Formulate the state-space representation of the system with the output variables:
i. The current from the source.

Flywheel
Rotational Inertia JL

Bearings
Damping b

Electric Motor
Transduction KT
Rotational Inertia J M
Electrical Resistance R
Voltage Source v(t)

Problems

629

Rigid Attachement

Angular Velocity Source, (t)

Compliant Shaft K
Gear
N2 teeth

Shaft rotates in
ideal bearing

Compliant Shaft
Torsional Spring K
Pinion, N1 teeth

Pinion
N1 teeth

Rack, N2 teeth
per inch
Mass M

(t)
Fluid coupling
Damping b

Inertia J
Gear
N3 teeth

Fig. P10.6 Rotational mechanical system

ii. The back EMF.


iii. The angular velocity of mass moment of
inertia,JL.
10.5.e Solve the state and output equations of part d for the
sinusoidal inputs of part c for the duration, t 10 ,
1
where =
.

min

Problem 10.6 The rotational mechanical system shown in


Fig. P10.6 consists of angular velocity source, (t), driving a fluid coupling with damping, b = 80 N m sec /rad .
The output shaft of the fluid coupling drives a pinion with
N1 teeth. The pinion with N1 = 10 teeth is meshed with
Gear Two with N 2 = 20 teeth and Gear Three with N 3 = 30
teeth. Gear Two is mounted on a compliant shaft with spring
constant, K = 600 N m/rad . The other end of the compliant
shaft is attached rigidly to the machine frame. Gear Three
is mounted on a rigid shaft which carries rotational inertia,
J = 6 kg m 2 , on an ideal, frictionless bearing.
10.6.a Derive the system equation for:
i The angular velocity of inertia, J.
ii The torque acting through the compliant shaft.
iii The torque applied by the angular velocity source.
10.6.b Derive the transfer function for the system equation
of part a and plot the poles and zeros (if any) of transfer function on the s-plane.
10.6.c Determine the steady-state response of the transfer
functions of part b for:
rad
sin(0.1t )
i (t ) = 10
sec
rad
ii (t ) = 10
sin(0.5t )
sec
rad
sin(t )
iii (t ) = 10

sec
10.6.d Formulate the state-space representation of the system with the output variables of part a.

Lubricant Film
Damping b
Fig. P10.7 Hybrid rotational-translational mechanical system

10.6.e Solve the state and output equations of part d for the
sinusoidal inputs of part c for the duration, t 10 ,
1
.
where =

min

Problem 10.7 The rotational-translational mechanical system, shown schematically in Fig. P10.7, consists of an angular velocity source, (t), acting on a compliant shaft modeled
as a torsion spring with spring constant, K = 120 N m/rad .
The spring drives a pinion with N1 = 12 teeth, engaged in a
rack with N 2 = 3 teeth per inch. The rack is attached to mass,
M = 5 kg . The rack and mass both slide on a lubricant film
with damping, b = 3 N m sec /rad .
10.7.a Derive the system equation for:
i The torque acting through the compliant shaft, K.
ii The velocity of mass, M.
iii The force acting to accelerate mass M.
10.7.b Derive the transfer function for the system equation
of part a and plot the poles and zeros (if any) of transfer function on the s-plane.
10.7.c Determine the steady-state response of the transfer
functions of part b for:
rad
sin(24t )
i (t ) = 10
sec
rad
ii (t ) = 10
sin(48t )
sec
rad
iii (t ) = 10
sin(96t )
sec
10.7.d Formulate the state-space representation of the system with the output variables of part a.
10.7.e Solve the state and output equations of part d for the
sinusoidal inputs of part c for the duration, t 10 ,
1 .
where =

min

10 Frequency Response

630

Problem 10.8 Plot the log-magnitude and phase-angle Bode


plots, using MATLAB or Mathcad, for transfer functions:
50
i G ( s ) =
s ( s + 25)
11
ii G ( s ) =
s ( s + 3)( s + 16)
320
iii G ( s ) =
( s + 4)( s 2 + 3s + 55)
4 s ( s + 10)
iv G ( s ) =
( s + 20)(4 s 2 + 8s + 64)
1
v G ( s ) =
2
( s + 1) ( s 2 + s + 2)
Problem 10.9 Plot the log-magnitude and phase-angle Bode
plots, using MATLAB or Mathcad, for transfer functions:
i

28(8s 2 + 96 s + 1, 280)
G (s) =
s ( s 2 + 112 s + 262)

239 s 2 + 5497 s
ii G ( s ) =
3
0.5s + 11.8s 2 + 386 s + 1260
8(60 s + 7800 s )
s 3 + 320 s 2 + 19, 900 s + 168, 000

iv G ( s ) =

( s + 3)
( s + 5s + 4)( s 2 + 4 s + 8)

v G ( s ) =

8(3s 2 + 0.6 s + 3)
s (2 s + 50)( s 2 + 20 s + 1200)

Problem 10.10 For transfer functions below:


10.10.a Sketch the log-magnitude vs. frequency plot using
the asymptotic approximation.
10.10.b Sketch the phase angle vs. frequency plot.
50
i G ( s ) =
s ( s + 25)
11
ii G ( s ) =
s ( s + 3)( s + 16)
320
iii G ( s ) =
( s + 4)( s 2 + 3s + 55)
4 s ( s + 10)
iv G ( s ) =
( s + 20)(4 s 2 + 8s + 64)
1
v G ( s ) =
( s + 1) 2 ( s 2 + s + 2)
Problem 10.11 For transfer functions below:
10.11.a Sketch the log-magnitude vs. frequency plot using
the asymptotic approximation.
10.11.b Sketch the phase angle vs. frequency plots.
G (s) =

ii G ( s ) =

v G ( s ) =

8(3s 2 + 0.6 s + 3)
s (2 s + 50)( s 2 + 20 s + 1200)

Problem 10.12 Plot the Nyquist polar plot, using MATLAB


or Mathcad, for the transfer functions:
i

G (s) =

ii G ( s ) =
iii G ( s ) =
iv G ( s ) =
v G ( s ) =

iii G ( s ) =

8(60 s 2 + 7800 s )
s 3 + 320 s 2 + 19, 900 s + 168, 000
( s + 3)
iv G ( s ) = 2
( s + 5s + 4)( s 2 + 4 s + 8)
iii G ( s ) =

50
s ( s + 25)
11
s ( s + 3)( s + 16)
320
( s + 4)( s 2 + 3s + 55)
4 s ( s + 10)
( s + 20)(4 s 2 + 8s + 64)
1
( s + 1) 2 ( s 2 + s + 2)

Problem 10.13 Plot the Nyquist polar plot, using MATLAB


or Mathcad for the transfer functions:
i

G (s) =

ii G ( s ) =

28(8s 2 + 96 s + 1, 280)
s ( s 2 + 112 s + 262)
239 s 2 + 5497 s
0.5s + 11.8s 2 + 386 s + 1260
3

8(60 s 2 + 7800 s )
s 3 + 320 s 2 + 19, 900 s + 168, 000
( s + 3)
iv G ( s ) = 2
( s + 5s + 4)( s 2 + 4 s + 8)
iii G ( s ) =

G (s) =

8(3s 2 + 0.6 s + 3)
s (2 s + 50)( s 2 + 20 s + 1200)

Problem 10.14Determine the transfer function of the


system, which produced the Bode plots in Fig. P10.14.
Problem 10.15Determine the transfer function of the
system, which produced the Bode plots in Fig. P10.15.
Problem 10.16 Determine the transfer function of the
system, which produced the Bode plots in Fig. P10.16.
Problem 10.17Estimate the transfer function that produced
the Nyquist plot in Fig. P10.17.

28(8s 2 + 96 s + 1, 280)
s ( s 2 + 112 s + 262)

Chapter 10 Appendix

239 s 2 + 5497 s
0.5s + 11.8s 2 + 386 s + 1260

Drawing Bode Plots in Mathcad

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


unknown system

120

80

40

log|G( j)|

dB
0

-2

-40

-4
0.01

-45

0.1

10

100

-80
1,000

0.1

10

100

1,000

rad
Input Frequency , ___
sec

-90o

o
G( j) -135

-180

-225 o
0.01

two orders of magnitude below the lowest corner frequency


to two orders of magnitude above the highest corner frequency. For example, if the systems corner frequencies were
0.8rad/sec and 70rad/sec, two orders of magnitude below
is approximately 0.01rad/sec and two orders of magnitude
above is approximately 10,000rad/sec.
The values of frequency within the range for which the
log-magnitude and phase-angle plots are calculated should
be spaced logarithmically to produce smooth curves. One
thousand points is more than enough for a smooth curve.
Calculate the frequency values as shown in the Mathcad
code below:
j := 1
low := 2
high := 3
p := 0..1000

w p := 10

p
low + ( high low)
1000

rad
Input Frequency , ___
sec

Where low is the logarithm of the lower limit of the frequency range and high is the logarithm of the upper limit of the
frequency range. Although Mathcad has the imaginary number i, the syntax to multiply by the imaginary number it must
be written as 1i. It is easiest to define j.
Log-magnitude plots are log G ( j w p ) using the magnitude bars created by typing the single vertical bar above
the forward slash. Be careful not to select the determinant
operator from the matrix pallet which has the same symbol
of two vertical bars. The code to calculate the log-magnitude of the transfer function can be in the place holder of
the ordinate (vertical axis) of the plot or the log-magnitude
can be calculated and assigned to a vector variable, say
LogMagp, before plotting. Remember to multiply the frequency by the imaginary number, j, to evaluate the transfer
function, G( j w p).
Calculate the phase angle, which is the angle or the argument of G( j w p ) with the arg() function. The arg() function returns the principle angle of a complex number

10 Frequency Response

632
Fig. P10.15 Bode plots of an
unknown system

20

-1

-20

dB

log|G( j)|
-2

-40

-3

-60

-4
0.01

45
0

G( j)

0.1

10

100

-80
1,000

0.1

10

100

1,000

rad
Input Frequency , ___
sec

-45

-90o
-135

-180

in radians. The principle angle of a complex number is


between G ( s ) It is convention to plot the phase
angle in degrees versus the excitation (input) frequency
in rad/sec, so multiply the argument by 180/ to yield
degrees. Right click within the plot to access the Format
menu. Format the frequency (horizontal) axis as a logarithmic scale.

Drawing Nyquist Plots in Mathcad


Nyquist plots are direct plots of the complex value of a transfer function evaluated substituting for the Laplace variable
s=j. Use Mathcads Re() and Im() functions to extract the
real and imaginary components from the complex number.
The Im() returns a purely real number with the correct sign.
The Nyquist plot is then produced as an xy plot with the xaxis Re(G ( j w p )) and the y-axis Im(G ( j w p )) .

0.01

rad
Input Frequency , ___
sec

Drawing Bode Plots in MATLAB


MATLABs Control Tool Box, an add-on package, contains
a Bode-plot function. What follows is the procedure for creating Bode plots without the Control Tool Box.
The MATLAB function for creating a semilogarithmic plot
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
xdata vectors provided by the user are not logarithms. The
function calculates the log(xdata) to create the plot.
The first step in creating a Bode plot is to determine the
desired input frequency range (spectrum). If in doubt, start
with a range which extends from approximately two orders
of magnitude below the lowest corner frequency to two orders of magnitude above the highest corner frequency. For
example, if the systems corner frequencies were 0.8rad/sec
and 70rad/sec, two orders of magnitude below is approximately 0.01rad/sec and two orders of magnitude above is
approximately 10,000rad/sec.

Chapter 10 Appendix

633

Fig. P10.16 Bode plots of an


unknown system

log|G( j)|

40

20

-1

-20

-2

G( j)

135

90

45

10

100

1,000

-40
10,000

10

100

1,000

10,000

rad
Input Frequency , ___
sec

0
-45

-90

-135

The values of frequency within the range for which the


log-magnitude and phase-angle plots are calculated should
be spaced logarithmically to produce smooth curves. One
thousand points is more than enough for a smooth curve. The
MATLAB code below spaces the frequencies at which values are calculated logarithmically, calculates the log-magnitude values, and creates the plot:
% Semilog plot. The Bode log magnitude
%of the transfer function: G(s)=100/(s^2+5*s+100)
j=sqrt(-1) % Define the imaginary number j
low = -2 % Exponent of the lower limit of freq range
high = 3 % Exponent of the upper limit of freq range
range = high - low % Number of decades in freq range
N = 1000 % Number of calculations in plot
w = zeros(N); % Preallocate and zero freq vector w
for p = 1:N;
% Calculate the freq value w(p);
w(p) = 10^(low+range*(p/N));

0.1

dB

0.1

rad
Input Frequency , ___
sec

% MATLAB's common log function is log10( );


logmag(p) = log10(abs(100/((j*w(p))^2+5*j*w(p)+100)));
end
semilogx(w,logmag) % Create semilog plot
grid on % Draw grid on the plot
Although MATLAB has the imaginary number defined as
i and j, it advises the user to use the syntax 1i for speed
and robustness. It is easiest to define j using the square root
function sqrt(). The magnitude of a complex number is calculated using MATLABs abs() function, where abs is short
for absolute value.
Calculate the phase angle, using MATLABs angle() function. The angle() function returns the principle angle of a
complex number in radians. The principle angle of a complex number is between G ( s ) . It is convention to
plot the phase angle in degrees versus the excitation (input)
frequency in rad/sec, so multiply the argument by 180/ to
yield degrees. The following MATLAB code calculates and
plots a phase-angle plot:

10 Frequency Response

634
Fig P10.17 Nyquist polar plot of
an unknown system

Im
-2

-1

rad
___
3 = 15 sec

Re

95
|1.

-21.8

o
-144 -j

|3.7

1|

rad
___
1 = 5 sec

-j2
-j3
-j4
rad -j5
___
2 = 10 sec

G(s) plane

-j6
% Semilog plot of the Phase Angle of
% G(s)=100/(s^2+5*s+100)
j=sqrt(-1) % Define the imaginary number j
low = -2 % Exponent of the lower limit of freq range
high = 3 % Exponent of the upper limit of freq range
range = high - low % Number of decades in freq range
N = 1000 % Number of calculations in plot
w = zeros(N); % Preallocate and zero freq vector
phi = zeros(N); % Preallocate and zero logmag vector
%Calculation loop
for p = 1:N;
w(p) = 10^(low+range*(p/N));
% MATLAB's angle() returns principle angle in rad;
% MATLAB has pi as a predefined constant;
phi(p) = (180/pi)*angle(100/((j*w(p))^2+5*j*w(p)+100));
end
semilogx(w,phi) % Create semilog plot
grid on % Draw grid on the plot

% Nyquist G(s)=100/(s^2+5*s+100)
j=sqrt(-1) % Define the imaginary number j
low = -2 % Exponent of the lower limit of freq range
high = 3 % Exponent of the upper limit of freq range
range = high - low % Number of decades in freq range
N = 1000 % Number of calculations in plot
RealG = zeros(N); % Preallocate and zero RealG vector
ImagG = zeros(N); % Preallocate and zero ImagG vector
%Calculation loop
for p = 1:N;
% Calculate value of freq w at a logarithmic spacing;
w = 10^(low+range*(p/N));
% Calculate value of G(j*w);
G = 100/((j*w)^2+j*w*5+100);
RealG(p) = real(G);
ImagG(p) = imag(G);
end
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 transfer function evaluated substituting for the Laplace variable
s=j. MATLABs real() and imag() functions extract the
real and imaginary components from a complex number. The
imag() function returns a purely real number with the correct
sign. The Nyquist plot is then produced as an xy plot with
the x-axis real (G ( j w p )) and the y-axis imag (G ( j w p )) .

Ogata K (2003) System dynamics, 4th edn. Prentice-Hall, Englewood


Cliffs
Ogata K (2009) Modern control engineering 5th edn. Prentice-Hall,
Englewood Cliffs
Rowell D, Wormley DN (1997) System dynamics: an introduction.
Prentice- Hall, Upper Saddle River
Shearer JL, Murphy AT, Richardson HH (1971) Introduction to system
dynamics. Addison-Wesley, Reading

AC Circuits and Motors

11

Abstract

Alternating current (AC), varying sinusoidally at 50Hz in Europe and Asia and 60Hz in
North America and Japan, is the dominant form of electric power. AC provides the continuously 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 deriving the transfer function of an electric circuit. Analysis of AC circuits is an application of
frequency response. Phasors are vectors in the complex resistancereactance 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.

11.1Introduction

11.1.1Alternating Current

Development of electrical science and technology began with


systems powered by voltaic piles or batteries which produced a constant voltage. It was natural that the first electrical
generators (or dynamos) produced constant voltage, yielding
what is termed direct or non-alternating current. Alternating current (AC) electric motors have powered stationary
machines for over 100 years, progressively replacing expensive steam engine driven shaft and belt systems. AC electric
motors are now used in combination with internal combustion engines in hybrid automobiles and have replaced internal
combustion engines in automobiles designed for commuters.
Mechanical engineers must understand AC motors and AC
circuits in order to use them effectively in their designs.
Alternating current circuits are driven by a sinusoidal
voltage source and are analyzed in steady-state. Consequently, the analysis of AC circuits, including AC motors, is an
application of frequency response analysis. We will begin
by introducing complex impedance, which is a method to
generalize the reduction of an electric circuit or linear graph,
which allows a graphical reduction of the energetic network
to a transfer function. We will then introduce some of the
terminology and technology of AC motors, including threephase AC power. Then we conclude with examples of
power calculations for three-phase electric motor.

The input to an AC circuit is always a sinusoidally varying


voltage v(t ) = V0 sin(w t ). AC circuit analysis is frequency response analysis. Consequently, AC analyses apply only to
the steady-state condition. If we need to know the transient
response of a system driven by AC power, we must use a
conventional system dynamics analysis.
By convention, the input voltage is assumed to have zero
phase shift and, when plotted as a vector, lies on positive real
axis. Usually, the output variable of interest is the current
drawn by the AC circuit. The phase angle of the current
drawn by a system, i (t ) = I 0 sin(w t + ), is measured relative
to the input voltage. Current phase leads or lags are created
by capacitive or inductive energy storage within a circuit.
For many of our analyses, we will find it convenient to represent the circuit being driven by the source as a single complex impedance load. Complex impedances will allow us to
work with phasors. Phasors are complex exponentials with
magnitudes which are RMS values defined below. The angle
of the phasor is the phase shift by itself, equivalent to setting time t = 0, at an arbitrary moment in steady-state. Consequently, AC circuit analyses are quasi-steady-state analyses.

K. A. Seeler, System Dynamics, DOI 10.1007/978-1-4614-9152-1_11, Springer Science+Business Media New York 2014

635

11 AC Circuits and Motors

636

11.1.2AC Power
Alternating current is the worldwide standard for the distribution of electric power. When Thomas Edison (18471931)
developed the first direct current (DC) power generation and
distribution system in 1882, it was known that the AC could
be transformed by magnetic induction to increase voltages
and lower currents for transmission, reducing resistive power
loss. AC power could be used to illuminate the carbon filament incandescent lamps, also developed by Edison. However, electric motors at the time were all DC motors. George
Westinghouse (18461914) created the first AC power
system in 1886 for lighting. It was not until Nikola Tesla
(18561943), who received the patent for the AC induction
motor in 1888, began working with Westinghouse that AC
power gained popularity. The long distance transmission of
hydroelectric power, three phase at 40Hz was demonstrated
in 1891, Frankfurt, Germany. Westinghouse and Tesla are
credited with the first large scale hydroelectric power plant,
which started operation at Niagara Falls in 1895. It generated
two-phase power at 25Hz in the USA.
The United States and Canada now use the frequency of
60Hz (cycles/sec) established by Tesla. Elsewhere, AC frequency is 50Hz with a few exceptions. Why we now use
60Hz and the rest of the world uses 50Hz is a good question. The higher frequency is more efficient. (One suggestion
why 50Hz is used outside of North America is that it is more
metric than 60Hz. The rationale for 60Hz was the need
to synchronize generating stations within a grid, which was
done twice a day using a master clock.) Some AC calculations are performed with frequency in Hertz, but others will
require angular frequency with units of rad/sec. One conversion that is helpful to know is 60Hz = 377rad/sec.
The residential service voltage in North America is
110120 VAC. It is a range and does vary with the load on a
system and the distance from the electrical distribution line.
Typically, the voltage of 115 VAC is assumed in an analysis. This voltage dates back to Edisons carbon filament
incandescent lamp. Edison generated 110 VDC to deliver
100 VDC, allowing for approximately 10% resistive loss in
transmission. Countries that use 50Hz also distribute power
at higher voltage, which compensates for the lower efficiency of that frequency. The European Union is currently implementing a dual voltage system of 240 and 400 VAC with the
lower voltage used for household appliances and lighting,
and the higher voltage for large motors.
Resistive energy loss during transmission of electric
power is a function of current flow, not voltage.


PLost = Ri 2

(11.1)

We shall see that AC power calculations must consider the


phase shift between current and voltage, which occurs when
there is energy storage in the system. For the time being,

we will consider purely resistive circuits. If a circuit has no


energy storage, then the electric power drawn by the circuit
is calculated by the product of current and voltage, P = vi .
Accordingly, an increased voltage allows a decreased current
to deliver the same power.
Recall the relationships between voltage, current, inductance L, and flux linkage from Sect.5.7.3,




di
dt

(5.45)

d
= v12
dt

(5.46)

= N = Li

(5.47)

v12 = L

AC powers sinusoidally varying current produces timevarying magnetic fluxes in coils and time-varying flux
linkages . Consequently, by magnetic transformation the
voltage of AC power can be stepped up for transmission
and stepped down for use, decreasing voltage and increasing current. Although it is possible to step up and step down
DC voltages of large power flows, it is more expensive. DC
electrical energy must be converted to mechanical energy by
a DC motor, which then drives a DC generator to produce
DC electrical power at a different voltage. Paired DC motors
and DC generators are expensive to purchase and to operate,
since they need maintenance. AC voltages can be stepped up
or down using transformers which are far less expensive to
build and require little maintenance.
A variety of voltages are used for industrial motors. Large
motors use voltages greater than 110120 VAC for a number
of reasons, including economy. The hazard posed by higher
voltages is not an issue because large industrial motors are
hardwired by electricians. The following are the typical voltages: 208 VAC, 230 VAC, 460 VAC, 575 VAC, and 2300
VAC. High voltage is above 600 VAC. The advantage
of high voltage is reduced current for the same power. An
induction motor using 2300 VAC draws only 49 Amps to
produce 200hp. In comparison, the same type of motor
designed to run at 230 VAC would need 490Amps, and a
lot more copper, to produce 200hp. Industrial motors also
use variable frequency AC current for speed control. AC is
supplied to the speed control circuitry at 60Hz. The AC is
converted to a direct current (DC), and then to the needed
frequency using circuitry known as an inverter.

11.1.3Root-Mean-Square (RMS) or Effective


Values
AC voltage measurements do not report the true amplitude
of the sinusoidal voltage or current. The value reported as
volts AC, abbreviated as VAC, represents the average or

11.1Introduction

637

V0

200

V RMS

100

v(t) 0
VDC

VDC 0

-V0

-200

-100

(n-1)T

nT

t, periods

(n+1)T

0.01

0.02

t, sec

0.03

0.04

Fig. 11.1
Root-mean-square voltage VRMS level plotted against
the sinusoidal voltage. The units VDC mean the true amplitude of the
voltage

Fig. 11.2 Three electrical phases, a, b, and c, of 60Hz (377rad/sec)


120 VAC power. Time t=0 is an arbitrary time after the system has
reached steady-state

effective voltage. The effective value corresponds to


the constant, or DC, voltage which would deliver the same
power to a resistance. Calculation of the effective value of
an AC voltage cannot be based on the average value of the
sine wave. The average value of a sinusoidal wave, such
as v(t ) = V0 sin(w t ), which has no vertical offset from zero,
Fig.11.1, is zero due to the symmetry of the sine wave.
Consequently, a straight average of v(t ) = V0 sin(w t ) over
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
the contribution of large numbers to the calculation. A way
of compensating is to take the square root of the average
of squared function. This yields the root-mean-square or
RMS value.
Expressed mathematically, the root-mean-square (RMS)
value of the v(t) is

11.1.4Sinusoidal and RMS Voltages

1
(v(t ))2 dt

T 0

VRMS =

(11.2)

or


VRMS =

2
1
V0 sin (w t )) dt
(

T 0

(11.3)

This integral is evaluated in the appendix of this chapter. The


result is the effective or root-mean-square (RMS) value of
v (t ) = V0 sin (w t )

VRMS =

V0
2

= 0.707 V0

(11.4)

The amplitude, or peak value, V0 of the sine can be calculated


from its root-mean-square value


V0 = 2VRMS = 1.41 VRMS

(11.5)

Voltages reported as VAC (volts AC) are root-mean-square


(RMS) values, and not the peak or true amplitude of the sinusoidal voltage. The terminology can be confusing because
the true amplitude of a sinusoidal voltage is reported as VDC
(volts DC) even though it is a sinusoid. VAC units of voltage
with RMS amplitude are always used in AC calculations.

11.1.5Three-Phase Alternating Current


Electric power is generated as three-phase alternating current, which consists of three sinusoidally varying voltages
spaced in time by phase angles that differ by 120 degrees, or
2 /3rad. Three-phase power is generated simultaneously in a
single generator. Each phase is generated and transmitted in a
separate circuit, the windings or coils of which are arranged in
the repeating pattern of a, b, c around the circumference of the
generators stator. Each of the three phases is transmitted on a
different electrical conductor or line. The three sinusoidally
varying voltages are known as phases due to their different
phase angles. The word phase in this context is similar to its use
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
of the circuits is arbitrarily chosen as phase a, Figs.11.2 and
11.3, it defines the beginning of the cycle. Then the other two
circuits sinusoids are seen as shifted in phase from phase a.
An individual line can be used as single-phase power
when its voltage is dropped across the electrical load to a
neutral conductor at ground potential. Alternatively two
or three of the lines can be used together as polyphase
power. Two phases are delivered to residential electrical services. Most small motors in home appliances or hand tools
run on single phase 120 VAC. Large motors, such as central
air conditioning compressors, and large resistive loads, such
as electric stoves and electric dryers, use the voltage drop

11 AC Circuits and Motors

638

A sin(t)

phase c

c = -240

-4
___
=
3

Fig. 11.4 Frequency response relationship

phase a

phase b

b = -120 o

-4
___
=
3

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.
Figure 11.2 shows three sinusoidal voltages, a(t), b(t),
and c(t). The peak voltages are V0=170 VDC. Since:
170 VDC
2

= 120 VAC

We are free to choose which of the three conductors carries


phase a and to define time t = 0 to be the beginning of phase
as cycle anytime once the system has reached steady-state.
In calculations, the voltage of phase a is always assumed to
have the phase angle a of zero. In Fig.11.2, the solid trace
has been named phase a.
(11.6)
a (t ) = V0 sin (w t )

Phase b peaks after or lags behind, phase a by one-third of a
cycle of 2rads, b = 2 /3, Fig. 11.3.
2

b (t ) = V0 sin w t

(11.7)

Phase c lags behind phase a by two-thirds of a cycle,

c = 4 /3 .

c (t ) = V0 sin w t

Fig. 11.3 Phasor diagram of three-phase AC voltage. A phasors


magnitude equals the RMS or effective magnitude of the AC voltage or
current. The phasors are drawn at their phase angle, corresponding to
the beginning of an AC cycle

where = G(j)

A|G(j) |sin(t+)

G(s)

(11.8)

11.2Frequency Response of Electric Circuits


The frequency response relationship, introduced in
Sect.10.2, Fig.11.4, applies to all types of energetic systems, including electric circuits, if (1) the system is linear

iL

iR

iC

g
Fig. 11.5 Sinusoidally excited RLC circuit. The sinusoid in the source
symbol denotes AC voltage and that the system is in steady-state

(or modeled as being linear), (2) the input is a sinusoid, and


(3) the system has reached steady-state. AC circuit analysis
meets these criteria.
We will introduce the specialized techniques developed
for analyzing AC circuits by applying the general purpose
method we have applied to other energetic systems. Our example is the RLC circuit of Chap.1, Fig.11.5, now excited
by AC voltage, as indicated by the sinusoid in the source
symbol. If it is powered directly from the grid or through
a transformer to step down the applied voltage, the input
frequency is 60Hz=377rad/sec. The output variable of interest is the current i flowing from the voltage source.
Following our standard analysis, we write the energetic
equations which describe the system.
Compatibility: v (t ) = v12 + v23 + v3 g
Continuity: i = iR
Elements: v12 = RiR

iR = iL
v23 = L

Energy: E sys = E L + EC

iL = iC
diL
dt

EL =

iC = C

1 2
LiL
2

dv3 g
dt
1
2

EC = Cv32g

Recall that our only use for the energy equations has been
to establish the initial condition of the output variable when
an input is applied to an energized system. The energy equations 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.
To use the frequency response relationship, we must
derive the transfer function for the input voltage and the
output variable, and the current from the voltage source.
Thus far, our procedure for deriving a transfer function has
been to (1) reduce the equation list to a differential system

11.3 Complex Impedance

639

equation, (2) perform the Laplace transformation on the


differential system equation, and (3) create the ratio needed
for the transfer function. We will now vary the sequence
of operations and perform the Laplace transformation on
energetic equations before we reduce the set of energetic
equations.
An unknown, time-domain variable transforms to an unknown Laplace-domain variable. We often omit indicating a
variable is in the time-domain, but we usually indicate that a
variable is in the Laplace-domain.

I s
()

V (s)

Cs
CsR + CsLs +

12

+ v23 + v3 g

Continuity:
R

I (s) = I R (s)

I R (s) = I L (s)

I L ( s ) = IC ( s )

L {v } = L { Ri }

Elements:

12

L {v } = L
23

L {i } = L
C

I (s)
Cs
=

=
2
V ( s ) CLs + CRs + 1
V (s)

V12 ( s ) = RI R ( s )

diL
L
V23 ( s ) = LsI L ( s )
dt
dv3 g
C
I C ( s ) = CsV3 g ( s )
dt

V ( s ) = V12 ( s ) + V23 ( s ) + V3g ( s )

V ( s ) = RI ( s ) + LsI ( s ) +

IC ( s )
Cs

(11.9)

I (s)
Cs

V ( s ) = R + Ls + I ( s )

Cs
Output ( s )
Input ( s )

I (s)

V ( s)

1
R + Ls +

1
Cs

C
s
CL
CR
1
s2 +
s+
CL
CL

1
s
L
=
R
1
V (s)
s2 + s +
L
CL
I (s)

(11.11)

Complex impedance is a generalized, frequency-dependent


resistance. The complex impedance of an electric circuit element is created by:
1. Transforming the elemental equation using the Laplace
transform.
2. Rearranging the equation into the ratio of voltage over
current. The quantity on the other side is the complex impedance.

V ( s ) = RI R ( s ) + LsI L ( s ) +

Cs
CRs + CLs 2 + 1

11.3Complex Impedance

Reduction: Input V(s), Output I(s)

V (s)

By performing the Laplace transformation before reducing


the set of energetic equations, we are able to obtain the transfer function without deriving the differential system equation. This is the essence of complex impedance.

V ( s ) = V12 ( s ) + V23 ( s ) + V3 g ( s )

L {i} = L {i }
L {i } = L {i }
L {i } = L {i }

I (s)

I (s)

Compatibility:

L {v(t )} = L {v

Cs
Cs

(11.10)

Clear fractions and write the denominator polynomial in


standard form by ordering the terms and clearing the coefficient of the highest-order term.

V (s)
Z Complex Impedance
I (s)

(11.12)

Capital Z is the symbol for complex impedance.


The form of the complex impedance of an electrical element is easy to remember once it is recognized as the same
form as electrical resistance.
The term complex refers to the Laplace transformation of the time-domain differential equations. The Laplace
variable, s = + jw , is a complex variable. The term impedance connotes a generalized resistance to power flow
through an element, since resistance impedes the flow of
current and power in a circuit.

11.3.1Complex Impedance of a Resistor


Illustrating with the elemental equation of a resistance
v (t ) = Ri (t )

L {v (t )} = L { Ri (t )}
V (s)
= R ZR
I (s)

V (s) = I (s) R
(11.13)

11 AC Circuits and Motors

640
Fig. 11.6a Linear graph of
RLC circuit shown schematically in Fig.11.5, b Linear graph
of the RLC circuit with complex
impedances in place of time
domain elements

The complex impedance of a resistor, ZR, equals its resistance R because the elemental equation is algebraic, not differential 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.

11.3.2Complex Impedance of a Capacitor


The elemental equations of energy storage elements are transformed from differential equations to algebraic equations.
The transformation of differentiation with respect to time
leads to the Laplace variable s being included in the complex
impedance. The complex impedance of a capacitor is:

L {i (t )} = L

dv (t )
C
I ( s ) = CsV ( s )
dt

V (s) 1
=
ZC
(11.14)
I ( s ) Cs

11.3.3Complex Impedance of an Inductor


Inductors store energy in the magnetic field built by their
current. The time dependency of the flow of energy into and
out of the magnetic field in the time-domain requires a differential elemental equation. Hence, the complex impedance
of an inductor also includes the Laplace variable s.


L {v (t )} = L

ZR

ZL

ZC

C
g

di (t )
L
V ( s ) = Ls I ( s )
dt
V (s)
= Ls Z L
(11.15)
I (s)

11.3.4Complex Admittance
Complex admittance is the inverse of complex impedance,
as conductance is the inverse of resistance. Complex impedances have the form of resistance. Complex admittance has
the inverse form. The symbol for complex admittance is Y.

V (s)
I (s) 1
1
= R ZR
= =
YR (11.16)
I (s)
V (s) R ZR

The complex admittances of a capacitance and an inductance


have the same form as the complex impedance of a resistance.



V (s) 1
I (s)
1
=
ZC
= Cs =
YC (11.17)
I ( s ) Cs
V (s)
ZC
V (s)
I (s)
1
1
= Ls Z L
=
=
YL (11.18)
I (s)
V ( s ) Ls Z L

Complex admittances are useful in reducing a system with a


number of parallel elements, as we shall see in Sect.11.3.6.

11.3.5Reduction of the RLC Circuit Using


Complex Impedances
We will illustrate the use of complex impedances as a simplification of the reduction of the RLC circuit of Sect.11.2. In
Sect.11.3.5, we will introduce an even more efficient graphical reduction method.
The first step is to replace the elements of the linear graph
of the RLC circuit with the corresponding complex impedance, Fig.11.6.
The key aspect of complex impedance is that all complex impedances can be manipulated as if there were
resistances in a time-domain resistive circuit. Figure11.6b
shows the complex impedances of resistance, inductance,
and capacitance in series. These three different types of
complex impedances can be manipulated as if they were all
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
complex impedances in series and in parallel also sum to the
equivalent complex impedance.

11.3.5.1Reduction of Impedances in Series


We derived the relationships to replace resistors in parallel
or in series with single equivalent resistances in Chap.5.
Recall that the equivalent element carries the same current

11.3 Complex Impedance


Fig. 11.7a Impedances in
series, b Single equivalent
impedance

641

Z1

Fig. 11.8a Impedances in


parallel, b Single equivalent
impedance

Z2

Z1
1

Z equiv

Z equiv
1

Z2
(through variable) flow and has the same voltage (across
variable) drop across it as the portion of the network it replaces.
The energetic equations for the complex impedances Z1
and Z2 in series, Fig.11.7a, are:
Continuity: I1 ( s ) = I 2 ( s ) I ( s )

time consuming, however, because the result must be placed


over a common denominator and inverted to be useful.
The energetic equations for the complex impedances Z1
and Z2 in parallel, Fig.11.8a, are:
Continuity: I1 ( s ) + I 2 ( s ) I ( s )
Compatibility: V12 ( s ) = V12 ( s )

Compatibility: V13 ( s ) = V12 ( s ) + V23 ( s )

Element Impedances:

ElementImpedances:

V12 ( s )
= Z1
I1 ( s )

V12 ( s )
V (s)
V23 ( s )
= Z1
= Z Equiv
= Z 2 13
I1 ( s )
I (s)
I2 (s)
Series
Reduction of two impedances in series to a single equivalent
impedance:

I1 ( s ) + I 2 ( s ) = I ( s )


Z Equiv I ( s ) = Z1 I ( s ) + Z 2 I ( s )

series

Z Series = Z1 + Z 2 + + Z n

parallel

Z parallel
(11.19)

This reduction is applicable to any complex impedance. It


can also be extended to any number of complex impedances
in series.
Complex impedances in series:


V12 ( s ) V12 ( s ) V12 ( s )


+
=
Z1
Z2
Z Equiv

1
1 V (s)
1
V12 ( s ) + = 12

Z Equiv
Z1 Z 2 Z Equiv
1

Z Equiv I ( s ) = ( Z1 + Z 2 ) I ( s )
Z Equiv = Z1 + Z 2

V12 ( s )
= Z Equiv
I (s)
Parallel

Reduction of impedances in parallel to a single equivalent


impedance:

V13 ( s ) = V12 ( s ) + V23 ( s )


Z REquiv I ( s ) = Z R1 I1 ( s ) + Z R2 I 2 ( s )

V12 ( s )
= Z2
I2 (s)

(11.20)

11.3.5.2Reduction of Impedances in Parallel


The reduction of complex impedances in parallel to a single
equivalent impedance is straightforward. It is generally more

1
1
+
Z1 Z 2

Z2
Z
Z + Z2
=
+ 1 = 1
Z1Z 2 Z1Z 2
Z1Z 2
Z parallel =

Z1Z 2
Z1 + Z 2

(11.21)

The reduction of two complex impedances in parallel can be


extended to any number of complex impedances in parallel.
Impedances in parallel sum as inverses to yield the inverse of
the equivalent impedance:


1
Z parallel

1
1
1
+
+ +
Z1 Z 2
Zn

(11.22)

1
to Z parallel . The result for
Z parallel
two elements in parallel is easy enough to invert. Inverting

The effort is in inverting

11 AC Circuits and Motors

642

ZR

ZL

ZC
g

Z D.P

The input variable of an AC circuit is always voltage,


Vsource(s). Notice that the driving point impedance is the inverse of the transfer function for the current from the source,
Isource(s).
I
(s) =
1
Cs
= source
Z D.P. Vsource ( s ) LCs 2 + RCs + 1

Fig. 11.9 Sinusoidally excited RLC circuit. The sinusoid in the source
symbol denotes AC voltage and that the system is in steady-state

When the denominator is expressed in standard form by


clearing the highest-order term of its coefficient,
1

the sum of three inverted impedances for three elements in


parallel is more involved.
1
Z parallel
1
Z parallel

1
1
1
+
+
Z1 Z 2 Z3

1
Z D.P.

Z 2 Z3
ZZ
ZZ
=
+ 1 3 + 1 2
Z1Z 2 Z3 Z1Z 2 Z3 Z1Z 2 Z3
=

Z 2 Z3 + Z1Z3 + Z1Z 2
Z1Z 2 Z3

Z parallel =

Z1Z 2 Z3
Z 2 Z3 + Z1Z3 + Z1Z 2

1
Z parallel

Vsource ( s )
I source ( s )

(11.23)

We will replace the three impedance in series, Fig.11.9a,


with a single equivalent impedance, Fig.11.9b, by use of
Eq.11.20. The equivalent impedance represents all of the
system, except the source. The equivalent impedance is the
driving point impedance, ZD.P.
Z R + Z L + ZC = Z D.P. R + Ls +

1
= Z D.P.
Cs

The effort is expressing the driving point impedance as a


proper ratio.
R


Cs
Cs 1
RCs + LCs 2 + 1
+ Ls
+
= Z D.P.
= Z D.P.
Cs
Cs Cs
Cs
Z D.P. =

Vsource ( s ) LCs 2 + RCs + 1


=
I source ( s )
Cs

I
(s) =
= source
Vsource ( s )

1
s
L
1
R
s2 + s +
L
LC

(11.25)

11.3.7Graphical Reduction of Networks


of Complex Impedances

When a system is reduced to a source and a single equivalent impedance, the equivalent impedance is called the driving point impedance, since it is the impedance the source
driving the system feels. A driving point impedance is
used to determine the power the source must provide to drive
the system. The definition of driving point impedance is
Z D.P. =

1
Cs
LC
1 LCs 2 + RCs + 1

LC

the result, Eq.11.25, equals the transfer function we derived


previously, Eq.11.11.

11.3.6Driving Point Impedance

Z D.P.

I source ( s )
=
=
Vsource ( s )

(11.24)

The reduction linear graph or a circuit in which the elemental equations are expressed as complex impedances can be
performed as a graphical technique. The method is to successively combine elements first those in parallel and then
those in series, replacing them with equivalent elements.
The process is repeated until the network is reduced to the
source element and a single, driving point, complex impedance. The graphical reduction to the driving point impedance
is quick and easy. The work is in the back-substitution and
clearing fractions. The graphical algorithm is illustrated in
Figs.11.10 and 11.11.

11.4Phasors and Phasor Operators


Phasors are vectors used in AC circuit calculations to
represent the variables, voltage and current. The phasor representation of voltage and current differs from the complex
exponential representation of those variables in just three aspects. First, AC analysis is steady-state. The magnitude of
the phasor is constant. Second, the magnitude used is the
RMS magnitude, not the true magnitude, again, because it
is an AC analysis. All AC calculations are in steady-state.
Second, only the phase angle is used as the angle. The time
varying angle, t, is not included. Consequently, a phasor
diagram shows the relative position of the variables at the

11.4 Phasors and Phasor Operators


Fig. 11.10a Initial circuit, b
Parallel impedances Z and Z
replaced by their equivalent impedance Zequiv, c Series impedances Zequiv and Z replaced by
their equivalent impedance, the
driving point impedance, ZD.P.

643

Z
Z

g
Fig.11.11a Initial circuit, b
Parallel impedances Z and Z
replaced by their equivalent impedance Zequiv, c Series impedances Z and Zequiv replaced by
their equivalent impedance, the
driving point impedance, ZD.P.

instant that the input sinusoid begins a new cycle, the angle
equals an integer multiple n2. The phasors are vectors of
voltage or current with RMS magnitude vector-like representations of complex exponential.

Z D.P

Z
g

Z
g

Z equiv

Z
g

Z D.P

Z equiv
g

reactance

Z L= jL
ZR = R

, resistance

11.4.1Reactance and Resistance


Phasor operators are vectors used to represent the complex
impedance or transfer function of the AC circuit. Representing complex impedances and complex admittances as
vectors allows graphical computations. To eliminate possible confusion with admittance, jY is not used to identify
the imaginary (vertical) axis of a phasor operators complex
plane. Unfortunately, the vertical imaginary axis is identified
as jX, which takes some getting used to, or alternatively, by
its units. The units of the imaginary axis are j and are called
reactance to sound similar but distinctive from the units of
the real axis, which are units of resistance, Fig.11.12. Although reactance appears as an imaginary resistance it is
a physical quantity. Physically, reactance represents energy
temporarily stored in inductance or capacitance and then returned 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
in AC circuit calculations.
A complex impedance or admittance is expressed as
a phasor operator by evaluating the impedance or admittance using s = jw , where is the frequency of the AC
circuit, 377rad/sec, and then plotting the result as a vector. Note that this is the same substitution used with the
frequency response relationship. Again, AC analysis is a
specialized frequency response analysis. Phasor operators
can represent the impedances of individual elements or the

ZC =

-j
1
___
____
=
jC C

Fig. 11.12 Phasor operators for the electrical properties of resistance,


capacitance, and inductance

reactance

, resistance

Z DP = 3 - 1,330 j
Fig. 11.13 Phasor operator for the driving point impedance of the RLC
circuit, Eq.11.24

driving point impedance of the entire system, Tables11.1


and 11.2.
The driving point impedance ZD.P., Eq.11.24, can be expressed as a phasor operator when its parameter values are
known, Fig.11.13. We will use R = 3 , L = 10 H, and
C = 2 F and the input frequency of w = 377 rad/sec.

11 AC Circuits and Motors

644

the flux linkages are related as:

Table 11.1 Impedances of electrical elements


Resistance

Capacitance

Inductance

V (s)
= R ZC
I (s)

V (s)
1
=
ZC
I ( s ) Cs

V (s)
= Ls Z L
I (s)

Table 11.2 Phasor operators of electrical elements


Resistance
Capacitance
Inductance
V ( jw )
= R ZC
I ( jw )

Z D.P. =

V ( jw )
1
j
=
=
ZC
I ( jw )
jCw Cw

V ( jw )
= jLw Z L
I ( jw )

Vsource ( s ) LCs + RCs + 1


=
Z D.P. =
I source ( s )
Cs
2

s2 +

R
1
Cs +
L
LC
1
s
L

rad
3
rad
1

+
j 377
j 377
+
sec
10 H
sec 2 F (10 H )
Z D.P. =
1
rad
j 377
10 H
sec
1
10
8
Vsource ( s ) 5 10 + j1.13 10 sec 2
=
Z D.P. =
1
I source ( s )
3.77 107
sec 2

Z D.P. =

Vsource ( s )
= 3 1,330 j.
I source ( s )

11.5Electrical Transformers
Electrical transformers consist of two or more coils of wire
wrapped around the same ferromagnetic core, Fig.11.14.
The same magnetic flux flows through both coils. By convention, the input coil is called the primary and the output
coil is called the secondary. Although schematics show the
primary and secondary coils at different locations on the ferromagnetic core, in practice both coils are wound together.
The black dots are the dot convention which mark transformers and schematics with the lead that are positive simultaneously. Note that the schematic also shows the AC source
in the positive half cycle.
The number of turns in the primary and secondary windings are Np and Ns, respectively. Because the same flux
passes through both coils, the flux linkage of one coil is related to the flux linage of the second coil. Using the definition of flux linkage:

= N
and writing the flux through the two coils as:

p
Np

p
Ns

s =

Ns
Np

Exciting a transformer with a time-varying alternating current produces a time-varying flux . Consequently, the flux
linkage is time varying and produces voltage across the
coils.
The relationship between the input and output voltages in
the transformer is
N
V34 = s V12
Np
The relationship between the currents through the primary
and secondary coils is calculated using energy conservation, neglecting the energy lost as heat in the transformer.
In steady-state, using the sign convention that power flow is
positive into the transformer,

Pp + Ps = 0 V12 i p = V34 is
N
N
V12 i p = s V12 is i primary = s is
Np
Np
Most engineering students are familiar with the use of transformers to step up and step down voltages in electric power
distribution systems. Transformers have other important applications. One is to provide a magnetic interface between
electrical subcircuits. A second is to allow impedance
matching in alternating current (AC) circuits.

11.5.1Model of a Transformer with a Resistive


and Capacitive Load
We will consider an RC circuit connected to the secondary
coil of a doorbell transformer, Fig.11.15.
How does one visualize the response of this circuit? First
analyze, or break the system into pieces, to create simpler
models. The effect of the primary side of the transformer is
to produce a sinusoidal voltage across the coil of the secondary side. If we are not interested in a power variable on the
primary side, we can replace the transformer with an equivalent AC voltage source.
We have simplified our model but we still have the question, how does one visualize the response of a circuit excited
by a sinusoidal input voltage? We begin with two quasistatic cases (quasi means resembling or sharing some attributes of the actual) by considering the input voltage at its
maximum and minimum values. We can interpret these two
cases as if they were driven by a DC source. The jargon for
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 impedance on the secondary side

1
AC Source

ip

+
2

1
AC Source

+
2

iR

iC

C
4=g

Fig. 11.16 Transformer replaced with the equivalent AC voltage


source driving the resistive and capacitive load on the secondary side

or has swung negative. They are represented by signs adjacent to the AC voltage source. The positive case follows
the dot convention, if the electrical schematic or transform
provides it. If we are only interested in the variables on the
secondary side, Figs.11.16 and 11.17, it will not matter if
we are in phase with 180o degrees or out of phase with the
primary side.

11.6Three-Phase Power
We conventionally mix angular units when working with
AC. We must use [w ] rad/sec and [ ] rad when we
evaluate transfer functions and complex impedances.
However, the phase angle is usually expressed in de-

Secondary Coil
N s turns

ip

is

R
5

Primary Coil
Np turns

Z load

Primary Coil
Np turns
Fig. 11.15 AC circuit with a
transformer with a resistive and
capacitive load on the secondary side

is

Secondary Coil
N s turns

grees to draw phasor diagrams, and also when we report


our results, since degrees are easier to visualize than are
radians.
Algebraic expressions of trigonometric functions can be
difficult to visualize. For example, consider the voltages differences a(t)b(t), b(t)c(t), and c(t)a(t).

(
)
(

a (t ) b (t ) = V0 sin (w t ) sin w t 120o


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

Note that the phase shifts would need to be converted to radians to evaluate these equations. Figure11.18 shows a(t),
b(t), and a(t)b(t). Note that the voltage a(t)b(t) has a
phase shift relative to both a(t) and b(t). The phase shift is
easy to see in a phasor diagram.
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
c(t)? Fig.11.18 shows the voltages of phases a and b and the
voltage difference between phases a and b.
The magnitude of a(t)b(t), b(t)c(t), and c(t)a(t)
may be a surprise. It is a common error to presume that the
amplitude is V0, but V0 is the magnitude of the voltage of a
phase relative to ground. Comparing the voltages a(t) and

11 AC Circuits and Motors

646
Fig. 11.17 AC circuit on the
secondary side with a transformer, a resistive and capacitive load. a AC source swung
positive, b AC source swung
negative

+
-

iR

iC

iR

iC

C
4=g

4=g
400

a-b

c-a

Im

b-c

200

-200
-400

b(0) = V0 e
0

0.01

0.02

t, sec

0.03

0.04

Fig. 11.18 Voltages of phases a and b and the difference in voltage


between phases a and b

350

a-b

200

VDC 0
-200
-350

-120

|V
0|

VDC 0

b
0

0.01

0.02

t, sec

0.03

0.04

Fig. 11.19 Voltages of phases a and b and the difference in voltage


between phases a and b

b(t) with their difference a(t)b(t), we see that the amplitude of the difference a(t)b(t) is greater for the simple reason that they are on opposite sides of the v=0 line for most
of a cycle. What is likely a surprise is that the difference
a(t)b(t) peaks before either a(t) or b(t), Fig.11.19.

11.6.1Line-to-Line Voltage
We wish to determine the magnitude of a(t)b(t), that is, the
magnitude sin (w t ) sin w t 120o . It is possible to use trigonometric identities and formulae but a vector construction
using complex exponentials and Eulers sine formula is easier.

j0
|V0 | a(0) = V0e

-j120

Re

Fig. 11.20 Complex exponential vectors which represent a(t) and b(t)
at t=0

The first step is to create complex exponentials which represent a(t) and b(t), Fig.11.20. The vectors must have the correct
amplitude and phase angle . Since we are evaluating steadystate relationships that must hold for any arbitrary time t, we
are free to choose the time t at which we will draw the vector
diagrams. Themost convenient time is t=0. Phase a(t) has a
phase angle of zero, so it plots on the positive real axis. Positive angles are counter-clockwise. Phase b(t) has a negative
phase angle since it lags, or peaks later in time than does a(t).
We next subtract the vectors to determine the magnitude of their difference. First, we multiply vector b(0) by
minus one which is equivalent to rotating the vector 180o,
Fig.11.21, and then create a vector sum a (0) + b(0) to perform 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
the sides of the triangle formed by a(0), -b(0), and a (0) b(0)
and we know the angle between a(0) and -b(0). There is surely
a sophisticated formula which we could hunt for in a handbook
of mathematics that would yield magnitude and phase angle of
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
line perpendicular to the real axis, Fig.11.23.
Calculate the magnitude of the vector a (0) b(0) with
the Pythagorean theorem.

a ( 0) b ( 0) =

( V (1 + cos (60 ))) + ( V sin (60 ))


0

11.7Physical Principles of Three-Phase AC Motors


o

= V0 e

j60

Im

|V

0|

(0)

|V0 | a(0) = V0e j0


0|

|V

b(0) = V0 e

-j120

|V0 | cos(60o)

Re

Fig. 11.23 Complex exponential vectors which represent a(t), b(t),


and a(t)b(t) at t=0

Fig. 11.21 Complex exponential vectors which represent a(t), b(t),


and b(t) at t=0

Im
0|

b(0

|V

(0)

|V0| sin(60o)

60

|V0 |

Re

|V0 |

-120

b(0

0|

-b(0) = -V0 e

|V

-j120

Im

647

|V0 |
a(0) = V0e

-b(0)= V0e j60

Re

j0 o

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
3 times the voltage of a single phase relative to ground.
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
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)
relative to a(0) is easily derived from the vector construction
using trigonometry.

ab = tan 1

V0

Fig. 11.22 Complex exponential vectors which represent a(t), b(t),


and a(t)b(t) at t=0

( ) = tan sin (60 )


(1 + cos ( 60 ))
(1 + cos (60 ))

V0 sin 60o

0.866
= 30o
1.5

ab = tan 1

Factoring V0 out of the radical.

a ( 0) b ( 0) =

( V ) (1 + cos (60 )) + (sin (60 ))


2

a ( 0) b ( 0) = V0

(1 + cos (60 )) + (sin (60 ))


o

11.7.1Rotating Magnetic Vector

Expanding the first term in the radical

(1 + cos (60 ))
o

( )

( )

= 1 + 2 cos 60o + cos 2 60o

Rearranging

a ( 0) b ( 0) = V0 1 + 2 cos 60o + cos 2 60o + sin 2 60o

( )

( )

( )

Using the trigonometric relationship from a unit circle


cos 2 ( ) + sin 2 ( ) = 1

a ( 0) b ( 0) = V0 1 + 2 cos 60o + 1

( )

yields the resulting line-to-line voltage.



a ( 0) b ( 0) = V0 3


11.7Physical Principles of Three-Phase


AC Motors

(11.26)

A magnetic vector is a vector in the direction of a magnetic field. If the magnetic field is symmetric, like a field of
a solenoid, the magnetic vector lies on the axis of symmetry.
The magnitude or length of the vector conveys the strength
of the field. In three-phase AC motors, a magnetic vector is
created by the windings (coil) of each phase. The windings
of three phases are on the stator, which is the stationary portion of a motor, and spaced one-third of the circumference
or 120 degrees apart around the axis of rotation of the rotor.
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 variation of current and the spacing of the three sets of windings
120 apart on the stator lead to the resultant of the vector sum
of the magnetic vectors of the three phases to have a constant magnitude and to rotate about the axis of the rotor once
every electrical cycle, Figs.11.24 and 11.25. The rotating

11 AC Circuits and Motors

648

t3

Im

ab |V0|

0|

|V

a(0

(0
-b

|V0| sin(60o)

t4

Re

|V0 | cos(60o)

stator

60

a
b

resultant magnetic vector drives the rotation of the motors


rotor, Fig.11.26.

t1

11.7.2Three-Phase Synchronous Motors

Fig. 11.25a One cycle of


three-phase voltage, b Positive
direction of magnetic vectors,
c Vector sums of the phases
magnetic vectors and time t1, t2,
t3, and t4

200

t1

t2

t3

11.7.3Three-Phase Induction Motors


Induction motors are the most common AC motor. The induction motor was invented concurrently by Nikola Tesla
and Galileo Ferraris (18471897). An induction motors
magnetic field is created by eddy currents induced in
the rotors conductive bars by the stators magnetic field,
Fig.11.28. The requisite for induction of the eddy currents
by the stators magnetic field is for that field to rotate faster
than the rotor. The difference is rotational speeds of the stators magnetic field and the rotor is termed slip.

t4

t1

0.004

0.008

t, sec

0.012

Resultant

t2

Resultant

c(t)
- 200

stator

Fig. 11.26 Resultants of the vector sums of the phases magnetic vectors and times t1, t2, t3, and t4. The resultant vector represents the stators
rotating magnetic field

a(t)
b(t)

stator
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 of a(t)b(t)

The vector sum of the three stator phases is a vector of constant 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
rings or generated on the rotor by an exciter. Since the
rotors magnetic vector is created by a DC current, it is fixed
in its position relative to the rotor, Fig.11.27. The rotors
magnetic vector chases the stators rotating magnetic vector.
Synchronous motors must be in synchrony or they lose
torque. Consequently, synchronous motors are not self-starting. They must be started and brought up to synchronous
speed by another motor, or by an auxiliary set of windings
on the same rotor.

stator

0.016

Phase b
Phase a

c
Resultant

Phase c
Phase Windings Positive
Magnetic Vector Directions

t3

c
Resultant

t4

b
a

11.8 Three-Phase AC Circuits

649

Fig. 11.27 A schematic crosssection of a three-phase synchronous motor

Three phase AC energizes the


three separate field windings,
phases a, b, and c. The magnetic
vectors of the three phases sum to
create a magnetic vector which
rotates around the axis of
the stator at the AC frequency.

DC current creates a magnetic


vector fixed in position relative
to the rotor. The rotor rotates to
align its magnetic vector with the
stators magnetic vector.

Copper or
aluminum bars

Induced eddy current

The bars of the squirrel cage are shorted


or electrically connected by the end caps
which also provide mechanical support.
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
in a pet rodents cage

Percent of Rated Torque

a
b

rotor

11.7.4Variable Frequency Motors


Variable frequency motors are induction motors driven various speeds by varying the AC frequency. The variable frequency sinusoidal voltage is created by the motors driver,
which are the motors electronics and power supply. A variable frequency motor driver approximates a sinusoidal voltage from rectangular pulses, a process known as inverting,
since half of the sine wave is negative.
The idea of a variable frequency motor dates to the nineteenth century but was not practical until electronics were
developed and not economical until the electronics became
inexpensive, as it has in the recent decades.

11.8Three-Phase AC Circuits

400

Design A

Design D

300

11.8.1Wye (Y) and Delta () Three-Phase


Connections

200

Design C

100
0

Design B
0

20

40

60

80

Percent of Rated Speed

100

Fig. 11.29 NEMA standards specify four different induction motors


designs by their torque-speed curves

Induction motors are self-starting. They have no brushes,


since there is no need to conduct current to the rotor, and
have lower maintenance cost than motors with brushes.
NEMA standards specify designs of induction motors,
categorized by the shape of the torque-speed curve, Fig.11.29.

We will now more formally consider three-phase electrical current using phasors. Three phase lines can be connected in a Y (also spelled as wye) connection or a
delta connection, Fig.11.30. The names refer to the configuration of the connections, shown below, where the lines
are phasors which represent the voltage of each of the three
phases. The voltage drop in the wye connection is between
one phase and neutral, which is a common return line used
to complete each of the three circuits. Consequently, a wye
connection requires four lines. The voltage drop in the delta
connection is between two of the phases. The vector calculation we performed above found the line-to-line voltage
between phases a and b. This is a delta connection. We determined that the voltage drop between the two phases was
larger than the voltage of either phase to neutral by a factor
of 3. The delta connection needs only three lines and has
a larger voltage drop across a load than does the wye connection for the same phase voltages. If the delta connection

11 AC Circuits and Motors

650

a b

b
b
neutral

a
c

c
Fig. 11.30a Y or wye connection. The center terminal is neutral
which is at ground voltage. b Delta connection
Fig. 11.31 A three-phase wye
connection generator

n
-

AC Source

+
c
needs fewer wires and provides a larger voltage drop, why
use a wye connection to drive a three phase motor? Many
three-phase motors in fact use both types of connections to
maximize performance by varying torque or power, or to
provide two speeds, where a wye connection is used for low
speed and a delta connection is used for high speeds. Motors operated this way are called single winding-two speed
motors.

11.8.2Wye Connected AC Machines


A wye connection is easier to understand and handle analytically. A schematic of a three-phase wye connected generator is shown Fig.11.31. The three AC sources, a, b, and c,
shown in the schematic are in fact three sets of armature
windings within one generator. The a, b, and c windings are
spaced around the stationary frame of the generator, called
the stator, such that when the generators rotor carrying
the field winding (a DC electromagnet) passes over the
armature windings in succession, the voltages induced in the
three windings peak at successively later time. This yields
three sinusoids which have phase shifts of 120.

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
mean if AC voltage is a sinusoid symmetric about zero? The
polarity of an AC source refers to the phase shift. Reversing
the source polarity it is equivalent to inverting the sine wave,
which is equivalent to a phase shift of 180.
The neutral line is needed to complete each of the three circuits. If you have done any home wiring, you know that neutral and ground are the same terminals in a residential electric
service. Grounding the neutral line assigns it a voltage of zero.
If we consider just one phase, we can visualize the positive
terminal of the phase swinging through the sinusoid voltage,
alternately pushing current into the neutral line and pulling
out of the neutral line. As it happens, when the three phases
are connected to the same neutral line, if they have a balanced load such as a three-phase motor which pulls the same
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.
However, it must be present for safety. It will carry current in
the event the load becomes unbalanced.
It is conventional to analyze three-phase circuits by considering just one of the phases and then tripling the resulting
current and power calculated. For example, a three-phase
generator driving a three-phase motor is shown in the schematic in Fig.11.32.
Any of the three phases can be considered in isolation,
but it is convenient to work with phase a since the voltage
supplied by phase a has a phase shift of zero. The schematic
becomes a familiar first-order RL circuit when it is redrawn
for any one of the three phases, Fig.11.33.
An alternative way to draw the wye generatorwye motor
schematic is to rearrange it so that it resembles a radiation
warning symbol, Fig.11.34.

11.8.3Delta Connected Machines


Delta connected machines do not have a neutral line,
Fig.11.35. The voltage drop through the windings is from
one phase to another. The voltage drop is greater than the difference between the voltage of one phase and ground when
one of the two phases is negative.
It is difficult to visualize the equivalent circuit of a single
phase in delta connections since the three phases are referenced to each other rather than to neutral as they are in the
wye connection. Fortunately, it is easy to convert a schematic
from a delta connection to a wye connection for analysis. In
fact, we developed the method when we calculated the line-

11.8 Three-Phase AC Circuits

651

Fig. 11.32 Three-phase generator and motor in wye connections

n
-

+a

AC Source

+
c
b

Fig. 11.33 Phase b windings of


a generator and a motor in wye
connections

c
Phase b

n
b

+ a
n

+
c

Fig. 11.34 A three-phase generator and motor in wye connection

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.
vab = van 3 = vbn 3
We must know if a reported terminal voltage is line-toline, vab, or line-to-neutral, van, in order to create the equivalent single-phase circuit. In industry, the term line voltage
means line-to-line. The term phase voltage means line-toneutral.
vline = v phase 3

neutral

Phase b windings
of AC motor

11.8.4Example 1 Three-Phase Delta Connected


Motor Phase Voltage

n
-

Phase b windings
of AC generator

The phase voltage of a delta connected motor is 460 VAC,


Fig.11.36. Determine the equivalent wye connected motor
for analysis.
We are given the phase voltage, or line-to-neutral, is 460
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.
vline = v phase 3 =

( 3 ) 460 VAC = 798 VAC

The equivalent wye connected motor is as shown in Fig.11.37.

11.8.5Example 2 Three-Phase Delta Connected


Motor Line Voltage
The line voltage of a delta connected motor is 460 VAC,
Fig.11.38. Determine the equivalent wye connected motor
for analysis.
In this case, the applied voltage of 460 VAC is reported as
line-to-line. The voltage which drops across the windings of
a motor in a delta connection is the line-to-line voltage. We
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.

11 AC Circuits and Motors

652
Fig. 11.35 Three-phase generator and motor in delta connections

+
-

+
-

Delta Connected
Three Phase AC Generator

460 VAC

460 VAC

Delta Connected
Three Phase AC Motor

460 VAC

460 VAC

460 VAC

Fig. 11.36 A three-phase motor in delta connection with 460 VAC


phase voltage

798 VAC

798 VAC

798 VAC

neutral

460 VAC

Fig. 11.38 A three-phase motor in a delta connection. Line voltage is voltage any between two of the three phases. If the reference
voltage(s) are not indicated, the default assumption is that voltages are
Line voltages

460 VAC

460 VAC

460 VAC

neutral

Fig. 11.37 Equivalent wye connected motor with line-to-neutral


voltages equivalent to the delta connected motor of Fig. 11.36

Fig. 11.39 Equivalent wye connected motor with line-to-neutral


voltages equivalent to the delta connected motor of Fig. 11.38

We do need to adjust the voltage applied to the equivalent


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
voltage drop, so we must increase the applied line-to-line
voltage to have the correct line-to-neutral voltage Fig.11.40.

11.9AC Power Calculation

vline = v phase 3 =

( 3 ) 460 VAC = 798 VAC

The product of current and voltage phasors is called apparent or total and has the units of voltsamperes.
Apparent Power Total Power V I
V I = V e j V I e j I = V I e j V e j I
j +
Apparent Power = V I e ( V I )

11.9 AC Power Calculation

798 VAC

653

798 VAC

798 VAC

neutral
Fig. 11.40 Equivalent wye connected motor with line-to-line voltages

jQ Reactive

Power, var

P
+j

Active Power
watts

jQ

Apparent
Power, VA

Fig. 11.41 Vector decomposition of apparent power. The two components of Total or Apparent Power vector projected on the real and
imaginary axes are active and reactive power

Since the AC source voltage has no phase shift, V = 0


Apparent Power V I = V I e

j I

[ Apparent Power ] = [ volts amps] = [ VA ] [ watts ]


Isnt a watt = i v ? It is for DC power but not for AC power.
DC power is the product of two real variables. Total power is
the product of a real and a complex variable or two complex
variables. When total power is expressed as a vector sum
the real component is the active power and the imaginary
component is the reactive power.
The real component P is active power. Active power is
power that is either dissipated as heat in the system or transduced from electrical to mechanical power in a motor, or
transformed from one AC current and voltage to another in
an electrical transformer. Excluding the amount dissipated as
heat, active power is what we can use as mechanical power.
Active power has units of watts (Fig.11.41).
The imaginary component of the apparent power vector, jQ, is called reactive power. Reactive power repre-

sents the flow of energy into either the magnetic field of an


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 magnetic 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 necessary for a motor to run, since an electric motors 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
the motor as mechanical power on the output shaft. Reactive power has units of var (lower case), which stands for
volts-amps-reactive.
Electric motors list a power factor on the patent
plate (name plate) and in the motors specifications. The
power factor is cos( ), where is the angle between the
total or apparent power and the positive real axis. Power
factors are often presented as 100 cos( ). The product of the
power factor and the apparent power vector is the projection of the apparent power vector onto the real axis, which
is the useful active power of the motor plus the power dissipated as heat.

11.9.1Example AC Power Calculations Using


Motor Specifications
11.9.1.1Example 1
A General Purpose single-phase AC 10hp motor specifications are shown below. Calculate the motors power.
Single-phase AC 10 hp motor specifications
Horsepower/Kilowatt
Voltage
Hertz
Phase
Full load amps
RPM
Service factor
Rating
Locked rotor code
NEMA design code
Insulation class
Full load efficiency
Power factor
Ship weight

10/7.46
230
60
1
41
1750
1.15
40C AMB-CONT
H
L
H
84
93
174 lbs.

We are given the voltage, the Full Load Amps (current), and
the motors power factor.
Calculate the apparent power.
Apparent Power = (230 VAC)(41A) = 9, 430 VA

11 AC Circuits and Motors

654

jQ Reactive
Power, var

jQ Reactive
Power, var
P
Active Power
watts

P = 8.7 kW
21.5
P+j

Q=9

jQ

.4 k

VA

Apparent
Power, VA

34

P+

jQ

Apparent Power

Power Factor
= Active Power
100

(9, 430 VA )(0.93) = 8, 770 W = Active Power


Reduce by the full load efficiency. Note that the Full
Load Efficiency is calculated after projecting the apparent
power vector onto the real axis.
Active Power Efficiency = Available Power

(8, 770 W )(0.84) = 7,370 W = Available Power


The motor is rated as 7.46kW, so our calculations are
accurate to the precision of the data.
To represent these results as phasors, we need the angle
between the apparent power vector and the real axis,
Fig.11.42.
power factor
1 93
o
= cos
= cos
= 21.5

100
100
1

11.9.1.2Example 2
General Purpose three-phase, Delta connected, 100 hp
motor specifications are shown below
Three-phase, Delta connected, 100 hp motor specifications
Horsepower/kilowatt
100/74.6
Voltage
230/460
Hertz
60
Phase
3
Full load amps
238/119
RPM
1185
Service factor
1.15
Rating
40C AMB-CONT
Locked rotor code
G
NEMA design code
B
Insulation class
B
Full load efficiency
94.1
Power factor
83
Ship weight
1267lbs

P
Active Power
watts

kV
A

Apparent
Power, VA

Fig. 11.42 Vector decomposition of apparent power

Calculate the active power. Multiplying the apparent power


by the power factor over 100 yields the projection of the apparent power vector on to the real axis.

=3

1.6

jQ

P = 26.2 kW

Fig. 11.43 Apparent power phasor and its projection on to the real,
active power, and imaginary, reactive power, axes

1. The Full Load Amps (current) given in a specification


of a three-phase motor is for one leg (or electrical supply line) of the three phases.
2. The rated voltage given in the specification (or on the
nameplate) of a three-phase motor is the line-to-line
voltage. Line-to-line voltage is measured between two of
the three phases, i.e., vab, vbc, vca. The voltage needed for
the power calculation is the line-to-neutral voltage. The
line-to-line voltage is a factor 3 greater than the line-toneutral voltage.
1
3

vrated = vline to neutral

Repeat the above calculations using the data for the threephase motor.
Calculate the apparent power for phase a, Fig.11.43.
1

Apparent Power = 230 VAC ( 238 A ) = 31.6 kVA


3

Calculate the power for phase a. Multiply the apparent power


by the power factor over 100 to yield the projection of the
active power vector on to the real axis.
Apparent Power

Power Factor
= Active Power
100

(31.6 kVA )(0.83) = 26.2 kW = Active Power


Reduce the active power of phase a by the full load efficiency to the shaft power.

( Active Power )( Efficiency ) = Shaft power of one phase


( Active Power )( Efficiency ) = ( 26.2 kW )(0.941)
Shaft power of one phase = 24.6 kW
The average shaft power from the other two phases is the
same. The average shaft powers over a cycle add as scalars,

11.10 Single-Phase AC Motors

655

so the shaft power is three times the average shaft power


from one phase.

Main Winding

L1

(3)( 24.6 kW ) = 73.8 kW


+

The rated power is 74.6kW. Our calculation is within


approximately 1%.
You will see the power calculations for three-phase motors performed using a formula which simplifies the calculation but removes some of the meaning.
Rated Power kW = 3 Rated VAC Full Load Amp
Power Factor Efficiency
To limit errors, it is important to see that the factor 3 results
from correcting from line-to-line voltage to line-to-neutral
voltage and from multiplying by three to sum the average
shaft power from all three of the phases.
3=

3
3

11.10Single-Phase AC Motors
A single stator winding excited by AC power can produce a
magnetic field which pulses in amplitude sinusoidally but cannot produce a rotating magnetic vector. The challenge in the
design of single-phase AC motors is to get them started. Once
there is rotation, then the interaction of the stator winding with
the rotor can create enough phase shift between the stators
and rotors magnetic vectors to keep the motor running.

Fig. 11.45 Cross-section of


a single-phase AC motor with
auxiliary starting windings

Starting
Winding

Centrifugal
Switch

L2
Squirrel Cage
Rotor

Fig. 11.44 Schematic of a capacitor start single-phase AC induction


motor

11.10.1Capacitor Start
A common design is a capacitor start single-phase AC
motor. A capacitor start motor has an auxiliary set of windings which is used to start the motor with the main windings,
Figs. 11.44 and 11.45. The starting windings are in series
with a large capacitor creating a phase shift between the starting windings and the main windings. The starting windings
are turned off by a centrifugal switch which opens when the
rotor has spun up to speed.

11.10.2Series Wound
A series wound single-phase AC motor is a high speed design.
The series connection between the stator and rotor windings
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.
Main Winding

Auxiliary Starting Winding

Rotor Winding

11 AC Circuits and Motors

656
Fig. 11.46 Cross-section of a
series wound single-phase AC
motor

Salient pole
Field winding
on the stator
The rotor and stator
(field) windings are
connected in series

11.11Mechanical Design Considerations


11.11.1NEMA
NEMA is a US electrical equipment standards body.
NEMA stands for National Electrical Manufacturers Association. NEMA standards exist for mechanical design
considerations, including motor performance, power supply requirements, frame sizes, mounting bolt patterns, electrical enclosure designs, and safety. The induction motor
classification based on the shape of its speedtorque curve
is a NEMA standard.

11.11.2US Department of Energy Efficiency


Data
Electric motors consume a large proportion of the electrical energy generated. Small improvement in the overall efficiency of the electric motors has both economic and environmental benefits. The US Department of Energy has
created a database of the specifications and efficiencies of
commercially available electric motors and offers free software to help engineers identify the most efficient motors for
particular design requirements.

Summary
AC power in North America is generated at 60Hz or
377rad/sec. Europe and most of the rest of the world use
50Hz, which is less efficient. Polyphase power is two phase
for residences and three phase for industry. AC circuit analysis is a specialized application of frequency response analysis. AC analyses are steady-state analyses. The magnitudes
of voltage and current are root-mean-square (RMS) values.
Conversion from the true, DC amplitude, V0, of sinusoidal
voltage to the RMS value is
VRMS =

V0
2

= 0.707 V0

stator
rotor

Complex impedances, Z, simplify the analyses and can be


used symbolically and graphically. The inverse of complex
impedance is complex admittance, Y. The dynamic attributes of an AC circuit or machine are commonly described
by its driving point impedance.
Phasors are vectors which represent the variables of AC
analyses, voltage and current. They are complex exponentials with RMS magnitudes and evaluated at time t=0, which
is set at the beginning of any cycle of the input voltage in
steady-state. Complex impedance represented as a vector
is termed a phasor operator. Phasor operators operate on
phasors. The complex plane of a phasor operator (complex
impedance) is the resistancereactance plane, where resistance is the real axis and reactance is the imaginary axis.
Reactance represents the energy temporarily stored in the
AC system during its cycle. The energy stored is returned to
the grid. The power is not used, but the electric utility must
provide the current.
AC powers complex planes real axis is active power.
The imaginary plane is reactive power. Power drawn as
shaft work from a motor is active power. The power factor of
a motor is 100 cos( ), where is angle between the phasor
operator and the positive real axis.

Problems
11.1 A single-phase AC motor with the driving point impedance Z = 0.5 + j 0.08 is connected to a 120 VAC
voltage source. Determine the current drawn by the
motor and the apparent, reactive, and active power of
the motor.
11.2 Two single-phase AC motors with impedances Z1 and
Z2 are connected in parallel to a 120 VAC voltage
source Fig. P11.2.

Determine the total current drawn by the motors and
the apparent, reactive, and active power of the each
motor.
11.3 A three-phase, delta connected AC motor with the driving point impedance Z = 3 + j1.2 is connected to a

Chapter 11 Appendix

657

Fig. P11.2a Motor circuit,


b Motor impedances

R2

L2

R1

L1

b
Z2 = 5 + 4.2 j

Z 1= 3.5 + 1.6 j

Fig. P11.5 An RL
AC circuit

R1

Fig. P11.7 An RL AC
circuit

R1

Fig. P11.6 An RC AC
circuit

R1

Fig. P11.8 An RLC AC


circuit

R1

R2

240 VAC voltage source. Determine the current drawn


by the motor and the apparent, reactive, and active
power of the motor.
11.4 A three-phase, wye connected AC motor with the driving point impedance Z = 3 + j1.2 is connected to a
240 VAC voltage source. Determine the current drawn
by the motor and the apparent, reactive, and active
power of themotor.
11.5 The RL circuit shown in Fig.P11.5 is powered by
120 VAC. Determine the driving point impedance for
R1 = 3 , R 2 = 20 , and L = 10 mH. Draw the phasor
operator of the driving point impedance. Determine the
current drawn by the circuit and the apparent, reactive,
and active power
11.6 The RC circuit shown in Fig.P11.6 is powered by 120
VAC. Use complex impedances to determine the driving point impedance of the RC circuit for R1 = 3 ,
R 2 = 30 , and C = 40 mF. Draw the phasor operator
of the driving point impedance. Determine the current
drawn by the circuit and the apparent, reactive, and
active power.

R2

R2

R2

11.7 The RL circuit shown in Fig.P11.7 is powered by


120 VAC. Use complex impedances to determine the
driving point impedance for R1 = 3 , R 2 = 20 ,
and L = 10 mH. Draw the phasor operator of the
driving point impedance. Determine the current
drawn by the circuit and the apparent, reactive, and
active power.
11.8 The RLC circuit shown in Fig.P11.8 is powered by
120 VAC. Use complex impedance to determine the
driving point impedance for R1 = 3 , R 2 = 20 , and
L = 10 mH. Draw the phasor operator. Determine the
current drawn by the circuit and the apparent, reactive,
and active power.

Chapter 11 Appendix
Evaluation of the Root-Mean-Squared Integral
This integral is evaluated twice, first in its present form and
then with the sine function expressed in terms of complex
exponentials using Eulers formulas.

11 AC Circuits and Motors

658
T

(V
0

We now evaluate integral expressed in terms of complex


e xponentials. Recall Eulers sine and cosine formulas.

sin (w t )) dt = V02 sin 2 (w t ) dt


2

Consulting a table of integrals for the integral of sine squared


yields,
1
1
sin (wt ) dt = 2 t 2w cos (wt ) sin (wt )
2

sin( ) =

1
1

cos (w t ) sin (w t )
V sin (w t ) dt = V t

2
2
w
0
2

e j + e j
2

e j (wt + ) + e j (wt + )
v(t ) = V0 cos(w t + ) = V0

Thus
T

cos( ) =

If we integrate over a full cycle, it makes no difference if we use cosine or sine. Using a cosine function
v(t ) = V0 cos(w t + ), express it using complex exponentials.

1
1
sin (wt ) dt = 2 t 4w sin ( 2wt )

2
0

e j e j
2j

1
2
V0 sin(w t + ) ) dt
(

T 0

VRMS =

2
0

or
T

1
1

sin ( 2 w t )
V02 sin 2 (w t ) dt = V02 t

w
2
4
0

2
2
t=
cos (w t ) t = 0w = cos w
cos ( 0) = 1 1 = 0

2
= sin w
sin ( 0) = 0 0 = 0

sin (w t ) t = 0w

VRMS

Either form will do since the integral is evaluated over a full


2
and
cycle of period T where T =

t=

sin ( 2w t ) t = 0w = sin ( 4 ) sin ( 0) = 0 0 = 0

V02
4T

VRMS =

(e

j (w t + )

+ e j (wt + )

(V
0

2
0 sin (w t )) dt = V0
2

1
T
t = V02
2 0
2

Substituting into the expression for root-mean-square,


VRMS =

2
1
V0 sin (w t )) dt =
(

T 0

VRMS =

V0

V
= 0
2 1.414

1 2 T
V0
2
T

(e

j (w t + )

+ e j (w t + )

) dt
2

) dt
2

= e j 2(wt + ) + 2e j (wt + ) e j (wt + ) + e j 2(w t + ) dt


0

(e

j (w t + )

+ e j (wt + )

(
(e

) dt = (e
2

j 2 (w t + )

+ 2e0 + e j 2(wt + ) dt

j wt + )

+ e j (wt + )

due to the symmetry of sine and cosine about zero. Hence

Expanding the integral

t=
2
sin ( 2w t ) t = 0w = sin 2 w
sin ( 0)

w
t=

T
e j (wt + ) + e j (wt + )
1
V0
=
dt

T 0
2

) dt = e
2

j 2 (w t + )

dt + 2 dt + e j 2(wt + ) dt

Evaluating the first integral term using eu


recognizing that the angular frequency w =
T

j 2
t +
T

dt =

e
0

j 2
t +
T

du deu
and
=
dx
dx

2
T

0 +
j 2
T j 2 T T +
T

e
e

dt =

T j ( 4 + 2 )
e j 2
e
4

Reference and Suggested Reading

659

The complex exponentials e j (4 + 2 ) and e j 2 are the same


complex number since 0, 2, 4, and any other integer multiple of 2 are the same angle. Therefore
T

j 2
t +
T

T j (4 + 2 )
T j 2
dt =
(e
e j 2 ) =
(e e j 2 ) = 0
4
4

and
T

j 2
t +
T

dt =

T j ( 4 + 2 )
e
e j 2
4

dt =

T j 2
e j 2 = 0
e
4

e
0

j 2
t +
T

leaving the only term in the integral to be


T

(e
0

j 2 (w t + )

+ 2e0 + e j 2(wt + ) dt = 2e0 dt = 2 (T 0) = 2T


0

Hence,
VRMS =
VRMS =

V02
4T

(e
0

j (w t + )

+ e j (wt + )

) dt
2

V02
V2
V
( 2T ) = 0 = 0
4T
2
2

Reference and Suggested Reading


Chapman SJ (2005) Electric machinery fundamental, 4th edn. McGrawHill, New York
Fitzgerald AE, Higginbotham DE, Grabel A (1981) Basic electrical
engineering. McGraw-Hill, New York
Fitzgerald AE, Kingsley C, Umans SD (2003) Electric machinery, 6th
edn. McGraw-Hill, NewYork

Index

A
AC induction motors,234
Across variable,123, 124, 126, 127, 130, 132, 135, 136, 147, 174,
181, 182
difference,125, 126
Active power,653
Address,495
Adjoint of a matrix,421
Alias,68
Alternating current (AC),635
Analysis,26, 31
Angular frequency,49, 69, 88
Angular velocity,222
Antilog,596
Apparent power,652
Area-moment of inertia,224
Attenuation,599
Automatic control theory,3, 4
Automotive lead-acid storage battery,47
B
Back-EMF,352, 354
Base conversion,496
Beat,586
Belt drive,339, 395
multiport,367, 384
Bending,200
Bevel gear,344
Binary,495
Binary place values,496
Bingham fluid,483
Bit,495
Block diagram,527
algebra,529
differential equation,535
state equations,542
Bode plot
critically damped phase-angle,603
first-order log-magnitude,598
first-order phase-angle,603
phase angle,602, 620
second-order log-magnitude,598
underdamped phase-angle,603
Bode plot asymptotic approximation,607
first-order factors,608
gain,614
integrator and differentiator,613
sketching,615
transfer function form,607
underdamped second-order factor,610

Bond graphs,123
Boolean control,3, 4
Boolean variables,499
Branches,19, 20, 24, 25, 35
Branch point,529
Brushes,351
Brushless DC motor,649
Byte,495
C
Cam and follower,362
Cantilevered beam,201
Capacitor start,655
Cartesian coordinates,65
Cartesian form,63, 66
Cascaded blocks,528
Cause and effect,148
Cavitation,347
Chain drive,339
Characteristic equation,54, 55, 57, 70, 71, 79, 545
Characteristic function,55, 57, 100, 159, 163, 164, 521, 524
Characteristic value see Eigenvalue,60
Check valve,483
Circular pitch,342
Closed-loop feedback control,3, 4
Closed-loop negative feedback control,531
Closed-loop system,533
Closed-loop transfer function,532
Cluster gear,346
Coast-down test,244
Coefficient matrix,522
Coenergy,12, 13, 14, 22
Cogging,351
Column matrix,412
Commutator,351
Compatibility,23, 28, 41, 125, 130
Compatibility equations,20, 21, 2326, 29, 31, 36, 37, 123
Complex exponential,60, 64, 67, 73, 83, 158
counter-rotating,71
form,6568, 80
imaginary component,70
real component,70
rotating,70
unit vector,64
Complex impedance,639
Complex number,56, 60, 61, 64, 65, 68
addition and subtraction,61
argument,63
cartesian form,60
complex conjugate,60

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

661

Index

662
magnitude,61, 63, 65, 83
modulus,61, 63, 83
multiplication and division,62
polar form,63
principal angle,64
Complex variable,54
Compliance,199
Compound gear,346, 388
Computational model,47
Continuity,21, 28, 32, 37, 125, 130
Continuity equations,2426, 29, 31, 36, 43, 123, 126, 134, 174, 182
Corner frequency,600
first-order,598
second-order,600
Controlled variable,531
Coulomb friction,211
torque on shaft,243
Crank and slider,362
Creep,235
Critically damped,88
D
DAlemberts force,122
Damped frequency,585
Damped natural frequency,549
Damping coefficient,120, 122, 157
Damping ratio,103, 550
Dashpot,217
DC generator,354
DC motor,350, 353, 354
DC servomotor,242
Decade,597
Decay,549
envelope,88
exponential,143, 147
rate,70
Decaying sinusoid,98
Decibel,596
Decision block,500
Decomposition,252
Delta connection,649
Design practice,27
Determinant,418
Diagonal matrix,420, 444
Differential elemental equation,140
Differential equations,45, 46, 57, 59, 75, 127
first order,142
Differential system equation,3, 16, 23, 37, 38, 42, 53, 93, 120, 121,
125, 127, 132, 135, 137, 139, 149, 182
Diode,483
Direct pass-through matrix,427
Disassemble,26
Discrete time,469
Disturbed system,54
Drag cup,232
Driving point impedance,642
Driving potential,20, 21, 23, 30, 35, 36
Dynamical systems,1, 2
electric circuits,3, 4
Dynamic calculation,70
Dynamic equilibrium,29
Dynamic range,86, 144
Dynamic system,2, 3, 4, 8, 31, 45, 46, 52, 135, 142
mathematics of,3, 5
Dynamic tests,235

E
Eddy currents,351
Effective mass,205
cantilervered beam,206
spring,206
Eigen equation,57
Eulers sine and cosine formula,68, 71, 73
Eigenvalue,57, 60, 69, 70, 75, 84, 155, 158, 179, 545
complex conjugate,60, 71, 79, 83, 88, 101, 159, 163
real,84
real component,89
underdamped systems,247
Eigenvector,444
Elastic beam theory,200
Elastic-perfectly plastic,5, 9
Electrical
capacitance,30
schematics,35
transformers,644
Element
damping,122
energetic,4, 20, 36, 118
energy storage,11, 30, 36, 127
equation,36
equations,126
ideal,27
linear,49
mass,118
mechanical,4, 7
Elemental models,32
Energetic
attribute,118, 135, 137, 174, 182
models,137
network,136
property,1, 8, 9, 20, 36
schematic,133, 154
systems,39
Energy,1, 8, 1315, 39, 157
density,14, 24
dissipation,118, 157, 168
elastic strain,10, 11, 32, 129
equations,37, 127
gravitational potential,15, 27
kinetic,15, 27, 121, 141
mechanical,39
potential,11, 16, 20
storage,30, 45, 46, 56, 135, 139
storage equations,127
strain,141
Engineering
analysis,5, 27, 39
computations,5, 8
design,26, 39
mathematics,3, 4
models,5, 39
Entangled loops,534
Equations
energetic,130
Equilibrium
dynamic,123
Equivalent element,255, 374
dampers in parallel,256
dampers in series,256
masses in parallel,258
springs in parallel,257
springs in series,257

Index
Error,531
Euler method,468, 469, 481
second-order system,471
Eulers equations,64
Eulers cosine formula,64
Eulers sine formula,65, 99
Eulers equations EulersEqs,65
Eulers number,143
Eulers sine formula,158
Executable block,500
Expansion by minors,419
Exponential,54
decay,55, 89
properties of,66
real,84
Exponentiation,596
F
Feedback,531
Feedforward loop,530
Feedforward transfer function,531
Feed-through matrix,522
Final value theorem,93
Finite difference,467
First order ordinary differential equation,74
First order step response,87
First-order step response,149
Floating,368
mechanical sources,368
transducer or transformer,373
Floating point variable,499
Flow chart,499
Fluid coupling,233
Flux,644
Flux linkage,636, 644
Force
source of,133
Force-source-mass-damper system,145
Force-source-spring-damper system,140, 151, 162
linear graph of,130, 132
Forcing function,57, 58, 79, 80
Forward stepping,468
Fourier series,46
Fourier transformation,49
Fourth-order RungeKutta
two state equations,479
Frequency response
equation,590
relationship,638
Fundamental frequency,586, 593
G
Gain,599
Gaussian elimination,96
Gear ratio,346
Gear set,334, 341, 384
multiport,365
Gear set,334, 341, 348
multiport,365
Gear train see Gear set,344
General solution,59, 76, 78, 82, 155, 157
Geometric compatibility,6, 32
Gibbs phenomenon,594
Graphical integration,17, 31
Graphical representation,27
Ground voltage,36

663
H
Harmonics,586, 593
Heaviside step
function,142
input transitions,182
Heaviside unit step function,46, 47, 49, 74, 76, 78, 91, 92, 136138,
150
superposition of,148
time shifted,148
Helical gear,343
Hexadecimal,495, 496
Homogeneous
equation,51, 54, 55, 59, 71, 75, 77, 79, 155
solution,55, 60, 75, 77, 79, 81, 82, 142, 144, 155, 157
Homogenous
response,51, 56
solution,58
Hookes law,29
Hybrid system
electrical, rotational, and translational,391
rotational translational,377
Hydraulic
actuator,356
motor,356
piston/cylinder,356
Hydrodynamic bearings,233
I
Ideal viscous friction,208
Identity matrix,97, 102, 523
Idler,384
Idler gear,346
Imaginary number,60, 62, 84
Impulse function,76
Impulse-momentum theorem,582
Impulse response
overdamped,247
underdamped,249
Incrementally linear,210, 521
Independent energy storage element,56
Inductance,636
Induction motors,648
Initial conditions,59, 76, 78, 80, 138, 140, 142, 155
energized systems,161
Initial value,146
method,165
theorem,93
Input
function,46, 52, 152, 153
impulse,138
pulse,148
pulse function,148
unit pulse,148
Input matrix,522
Input pulse function,151
Integral causality,594
Integrator,536
Invariant,24, 42
Inverse mathematical operations,144
Inverse of a matrix,523
Inverse tangent,83
algorithm,63
function,65
Inverted,529
Inverter,529

Index

664
J
Jenkins-Traub algorithm,57
K
Kinetic energy,144, 198
Kirkchoffs circuit laws,123
L
Lag,584, 587
Lagranges and Hamiltons energy methods,4, 7
Laplace-domain,90, 519, 521, 522, 526
response function,95
Laplace transform
pair,97, 103, 163
tables,95
Laplace transformation,3, 31, 38, 56, 71, 75, 90, 92, 93, 94, 99, 100,
155, 157, 163, 178, 521
initial condition terms,56
inverse Laplace transformation,92, 94, 95, 98, 100
inverse transformation,164
transform integral,90
transform pairs,95, 101
Laplace variable,54, 55, 90, 94, 95, 96, 521
Lead,587
Leading,584
Lenzs law,352
Lever,335, 338
multiport,363, 379
Linear algebra,96, 98, 101
Linear differential system equation,521
Linear elemental equations,183
Linear equations,149
Linear graph,121, 124, 135, 154, 181
drawing,218
nodes,217
Linear graph method,16, 30, 122, 129
Linear graph symbol
transformer,335, 337
Linearization,208
Linearize,53
Linear mathematical operators,50
Linear operator,53, 519, 527
Linear systems,3, 5
Line to line,651
Logarithmic
scale,597
spiral,70
Logarithms
common,595
properties,595
Log-decrement formula,249
Logical variables,499
Loop,22, 39
equation,23, 41, 126
M
Magnetic flux,352
Magnetic hysteresis,351
Magnetic reluctance,352
Magnetic vector,647
Magnitude,588
Magnitude ratio,581
Mapping,57
Mass,9, 14, 25, 197
conservation,15, 26
flow rate,20, 37
Mass-damper system,139, 161

Mass-moment of inertia,222
cylinder,223
parallel-axis theorem,224
superposition,224
Mathcad,106
assignment statement,106
Heaviside unit step function,190
plot,106, 190
superposition,190
Mathcads Heaviside step function,138
Mathematically independent, algebraic equations,21, 38
MATLAB
absolute value function abs,633
angle function,633
arithmetic operators,505
array,504
array and matrix operators,507
array variables,110
assignment statement,110, 508
Bode plots,632
colon operator,513
comments,510
control tool box,515
editor,110
elseif statement,509
else statement,509
environment,109, 503
flow chart,112, 193
for-end loop,110
for loop,509
formatting plots,511
function imag,634
function real,634
if statement,190, 508
logical operators,507
nested loops,191
Nyquist plots,634
plot,111
plot statement,510
plotting and figures,511
reading and writing to a text file,513
reading and writing to Excel worksheet,513
reading and writing to files,512
relational operators,191, 507
script,111, 112
semilogx,632
square root function sqrt.,633
time shift,191
user written functions,475
while loop,509
workspace,111
MATLAB code,470, 472, 474476, 478, 479, 481, 483
Matrix
addition,412
cofactors,418, 420
differentiation,414
inversion,416
multiplication,413
subtraction,412
Mechanical power,211
Mechanical system
rotational,20
translational,20
Mechanical transformers,334
Method of undetermined coefficients,46, 51, 74
Miter gears,344

Index
Model,27
abstract,5, 118
elastic-perfectly plastic,9, 50
element,29
energetic,8, 13, 118
engineering,2, 4, 7, 26, 27, 31, 32
graphical,28
input,137
linear,50, 53, 120
linear element,50
linear material,9, 15
lumped parameter,28
material,6, 10
mathematical,6, 7, 28, 84
network,20, 36
Newtonian,8, 13
nonlinear,7, 11
nonlinear stress-strain,7, 11
physical,31
Modulus,590
Motor action,350
Multiplicative operator,93, 522
N
Natural fractions,498
Natural frequency,249
undamped,103
Natural response,2, 3
NEMA,649
Nested loops,533
Network,19, 35
Network topology,24, 42
Newtonian fluid,119
Newtonian formulation,121, 122
Newtonian mechanics,14, 26
Newtons second law,15, 26
Node,10, 19, 20, 21, 25, 35, 38, 121, 124, 125, 132, 147
equation,125, 136
ground,132
subscript,125
subscripts,22, 39, 122
velocity,122
Non-causal,151
Non-inverted,529
Non-linear damping,209, 473
Nonlinear differential equation,53
Non-linearity
backlash,482
clipping,482
directional dependency,483
inflection point,482
saturation,482
threshold,483
Nonlinear physical relationship,51
Non-oscillatory see Overdamped,154
Normalization,144
Numerical differentiation,538
Numerical instability,90
Numerical methods,47
Nyquist polar plot,623
O
Observable canonical state-space form,539
Octal,459
Open-ended design problems,4, 6
Open-loop,532
Operating

665
condition,119
point,209
range,10, 16, 47, 209
Operator,527
Ordinary differential equation,53, 77
second order,79
Orienting,25, 44
Orthogonality,419
Oscillations,1, 16, 28, 56, 71
decaying of,69
period,88
Oscillatory
second order step response,88
second order system,99
Output matrix,522
Overdamped,86, 154
Overshoot,70
P
Parallelogram rule,61
Partial fraction expansion,92, 95, 560
Particular solution,51, 57, 78, 80, 142
Passive elements,124, 125
Path,19, 35, 125
equation,23, 41, 126, 136
equations,126
Period,581, 583, 593
Phase angle,69, 82, 84, 635
Phase shift,69, 81, 103
Phase voltage,651
Phasor operator,643
Phasors,642
Physical systems,58
Piece-wise continuous function,89
Pitch circle,342
Plant,532
Polar form,588
Pole,547
Pole-zero form,588
Pole-zero plot,551
Polyphase,637
Positive feedback,532
Post-multiplication,97
Potential,20, 36
Power,8, 12, 15, 16, 20, 130, 147
flow,15, 26, 124
ideal source,47
mechanical,12, 17, 21, 122
source,16, 25, 29, 46
variable,46
variables,1, 15, 16, 26, 143
Power law viscosity,474
Power screw,361
Power sources,212
rotational systems,234
Power variables,127
Practical pulse-width-modulated amplifiers,148
Pre-multiplication,97
Pressure,20, 21, 36
Proportional controller,533
Proportional damping,208
Pulse response,148
function,153
Pulse train,148
Pulse-width-modulation,148
Pump,354, 396

Index

666
axial flow,356
gear,355
positive displacement,354
reciprocating piston,355
turbopump,356
Pythagorean theorem,61, 62, 63, 81, 83
Q
Quadratic equation,77
R
Rack and pinion,360
Radix,496
Reactance,643
Reactive power,653
Real numbers,60
Recursive calculation,468
Reference signal,531
Relaxation,235
Relaxation time,53
Relevant physical truths,28
Resistive energy loss,636
Resonance,584
Resonant,586
Resonant frequency,585
Resonant peak,585
effect of damping ratio,601
Response
steady-state,143
Response function,89, 146, 172, 173
pulse,153
Reversible systems,16, 28
Rigid,12, 20
Rolling-contact bearings,232
Root-mean-square (RMS),637
Rotating magnetic field,648
Rotational damper,233
Rotational hydraulic motor
gear,359
gerotor,358
piston,359
rotor,358
vane,358
Rotational mechanical systems,221
Row matrix,412
Runge-Kutta algorithms,89
Runge-Kutta method,45
three state equations,480
S
Saint-Venants principle,9, 15
Schematic,124
Second order
spring-mass-damper system,156
underdamped unit step response,165
Second order step responses,86
non-oscillatory,86
overdamped,251
Second-order step responses
superposition of,154
Self-locking,362
Series wound,655
Serpentine belt,384
Shear modulus,230
Shear strain,231
rate,119
Shock absorber,210

Signal,527
Sign convention,16, 28
fluid transducers,355
linear graph,215
mechanical systems,214
motors,352
transducers and transformers,346
Significant,28
Signs,145
Simply supported beam,201
Singular matrix,420
Sink,214
Sinusoid,58
Sinusoidal inputs,49
Sinusoidal output,588
Sinusoidal response
transient,579
Slip,648
Slug,198
Snubbers,209
Solenoid,352
Source
force,122
power,122
Spectrum,598
Spring
translational,129
Spring back,14, 25
Spring constant,13, 23, 129, 199
Spring rate see Spring constant,13, 23
Spur gear,342
Square wave,594
Stability,547
Stable,84
exponential growth,76, 86
Standard form,53, 74, 77
State equations,522
standard vector-matrix form,522
vector-matrix form,426
State-space,423, 437
higher-order system equation,438, 440
State-space representation,149
State variables,139, 146, 147, 153
energy storage variables,544
without physical meaning,544
State vector,522
Steady-state,54, 58, 85, 147, 152, 161
response,59, 89, 142
value,146
Step function
Heavisides definition of,138
Step response,85
function,49
unit,151
Strain energy,199
Summation junction,529
Superposition,46, 49, 51, 52
Synchronous motor,648
Synthesis,26
System,19, 35
boundaries,45
energetic,20, 36
fluid,21, 38
identification,84
mass-damper,128
mechanical,12, 20

Index
nonlinear,151
physical,30
spring-damper,130
System dynamics,4, 6
System equation,121, 134
T
Thermal systems,30
Thermodynamics,16, 28
Three phase,637
Time constant,51, 53, 84, 85, 86, 89, 121, 142, 144, 146, 151, 152,
153
form,121
Time-domain block diagram,536
Time-domain expression,98
Time shift,18, 32, 147, 150, 153
Time step,89
Torque,222
Torque constant,352
Torque curve,234
Torsion spring,230
Transducer,336, 350
Transfer function,67, 93, 94, 99, 162, 521, 524, 527
Transformation matrix,443
Transformed state vector,523
Transformer,333
block model,336
Transformers and transducers
multiport,362
Transient period,51, 54, 161
Transient response
first-order system, sinusoidal input,582
Translational mechanical system,129
Translational spring,199
Translational velocity,129
Transpose of a matrix,415
Trapezoidal approximation,148
Trapezoidal integration,484
Trigonometric identities,71
Trivial equations,33
U
Underdamped,154
Underdamped pulse response
second order,165

667
Underdamped system,79
Unit exponential decay,85
Unit impulse function,46, 48
Unit ramp function,46, 48
Units,53, 128
Unit stable exponential growth,85
Unit step response,152
first order,153
function,46, 153, 178
Unstable,548
exponential growth,69
system,69
U.S. customary units,198
mass moment of inertia,226
V
Variable frequency motor,649
Vector algebra,411
Vector-matrix algebra,411
Vector of inputs,427
Vector of output variables,427
Vector of transfer functions,526
Velocity
source,129, 132, 212
Velocity source-damper-mass-spring system,174
Viscoelasticity,220
Viscoelastic model
KelvinVoight,220
Maxwell,220
Zener,220
Viscosity,119
W
Weight,197
Word,495
Work,8, 11, 12, 13
Worm drive,344
Wye connection,649
Y
Youngs modulus,9, 15
Z
Zero,551

Potrebbero piacerti anche