0% found this document useful (0 votes)
3 views

2_Automatic Control Systems and Components

The document is a textbook titled 'Automatic Control Systems and Components' by James R. Carstens, covering various aspects of automatic control systems, including electronic fundamentals, control theory, transducers, and system communications. It includes detailed chapters on mathematical concepts, design examples, and practical applications in process control systems. The book serves as a comprehensive resource for understanding and implementing automatic control systems in technology.

Uploaded by

Ashish Mahajan
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
3 views

2_Automatic Control Systems and Components

The document is a textbook titled 'Automatic Control Systems and Components' by James R. Carstens, covering various aspects of automatic control systems, including electronic fundamentals, control theory, transducers, and system communications. It includes detailed chapters on mathematical concepts, design examples, and practical applications in process control systems. The book serves as a comprehensive resource for understanding and implementing automatic control systems in technology.

Uploaded by

Ashish Mahajan
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 472

•J 1 ■

r#Tj

JAMES R. CARSTENS
Digitized by the Internet Archive
in 2018 with funding from
Kahle/Austin Foundation

https://archive.org/details/automaticcontrolOOOOcars
AUTOMATIC
CONTROL
SYSTEMS
AND COMPONENTS
AUTOMATIC
CONTROL
SYSTEMS
AND COMPONENTS

James R. Carstens, P.E.


ASSISTANT PROFESSOR
School of Technology
Michigan Technological University

PRENTICE HALL Englewood Cliffs, New Jersey 07632


Library of Congress Cataloging-in-Publication Data

Carstens, James R.
Automatic control systems and components / James R. Carstens.
p. cm.
ISBN 0-13-054297-0
1. automatic control. 2. Process control. I. Title.
TJ213.C293 1990
629.8—dc20 89-37977
CIP

Editorial/production supervision and


interior design: Diane M. Delaney
Cover design: Wanda Lubelska Design
Manufacturing buyer: Dave Dickey

© 1990 by Prentice-Hall, Inc.


A Division of Simon & Schuster
Englewood Cliffs, New Jersey 07632

All rights reserved. No part of this book may be


reproduced, in any form or by any means,
without permission in writing from the publisher.

Printed in the United States of America

10 987654321

ISBN []-13-0£45^7-0

Prentice-Hall International (UK) Limited, London


Prentice-Hall of Australia Pty. Limited, Sydney
Prentice-Hall Canada Inc., Toronto
Prentice-Hall Hispanoamericana, S.A., Mexico
Prentice-Hall of India Private Limited, New Delhi
Prentice-Hall of Japan, Inc., Tokyo
Simon & Schuster Asia Pte. Ltd., Singapore
Editora Prentice-Hall do Brasil, Ltda., Rio de Janeiro
To my wife, Sandy, and to my two children, Sig and Lesley,
all of whom have given meaning to my life.
Contents

Preface xvii

1: REVIEW OF ELECTRONIC FUNDAMENTALS 1

1-1 Introduction 1
1-2 The DC Circuit 1
1-2.1 The Ampere, 2
1-2.2 The Volt, 2
1-2.3 The Voltage Source, 3
1-2.4 Current Direction, 4
1-2.5 Determining Polarity, 4
1-2.6 Ohm’s Law, 5
1-2.7 Resistivity, 6

1-3 The Resistor 8


1-4 Series and Parallel Circuit Resistance and How
They Are Calculated 9
1-4.1 The Series Resistance Circuit, 10
1-4.2 The Parallel Resistance Circuit, 13
1-4.3 The Combined Series-Parallel
Resistance Circuit, 15
1-5 Kirchhoff’s Laws 17
1-5.1 Kirchhoff’s Voltage Law, 17
1-5.2 Kirchhoff’s Current Law, 18

VII
vm Contents

1-6 The Voltage Divider Rule 19

1-7 Power Supply Regulation 20


1-8 The AC Circuit 22
1-8.1 Generating an AC Signal, 22
1-8.2 Waveform Definitions and
Conventions, 23
1-8.3 The Phase Angle, 26

1-9 The Capacitor 28


1-9.1 Capacitor Characteristics, 28
1-9.2 Capacitors in Parallel, 30
1-9.3 Capacitors in Series, 31
1-9.4 Series-Parallel Capacitors, 31

1-10 The Inductor 31


1-10.1 Inductor Characteristics, 32
1-10.2 Inductors in Series, 33
1-10.3 Inductors in Parallel, 34
1-10.4 Series-Parallel Inductors, 34

1-11 The Transformer 34


1-11.1 Transformer Characteristics, 34
1-11.2 The Turns Ratio, 35
1-11.3 Power Transformer Ratings, 36

1-12 Transient Behavior 37


1-12.1 The R-C Circuit, 37
1-12.2 The R-L Circuit, 38

1-13 Reactance and Impedance 39


1-13.1 The Resistor’s Reactance to AC
Waveforms, 40
1-13.2 The Capacitor’s Reactance to an AC
Waveform, 40
1-13.3 The Inductor’s Reactance to an AC
Waveform, 42
1-13.4 What Is Impedance? 44
1-13.5 Understanding the J-Operator, 45
1-13.6 Instantaneous versus RMS Voltages
and Currents, 46

1-14 Power 48

1-15 Resonant Frequency 49

1-16 Oscillators 51
Contents ix

1-17 Solid State Devices 51

Summary 53

2: CONTROL THEORY BASICS 54

2-1 Frequency Response 54

2-2 Gain 56

2-3 Bandwidth 57

2-4 Linearity 57

2-5 Phase 57

2-6 Feedback 61

2-7 The Block Diagram 66

2-8 The Summing Point 67

2-9 The Transfer Function 68

2-10 An Automatic Control System 68


Summary 71
Exercises 72

LAPLACE TRANSFORMS

3-1 The Mathematics of an Automatic Control


System 74

3-2 A Little Background Concerning Laplace


Transforms 74
3-3 How Transforms Are Used 75

3-4 Transform Tables 78

3-5 Examples Using Transforms 81


Summary 83

Exercises 83

4: TRANSDUCERS AND CONTROL SYSTEM COMPONENTS 85

4-1 Defining Transducers 85


4-2 Position Sensing 87
x Contents

4-2.1 The Potentiometer, 87


4-2.2 The LVDT, 90
4-2.3 The Linear and Rotary Photocell
Encoder, 90

4-3 Velocity Sensing 91


4-3.1 The Rotary Magnetic Encoder, 93

4-4 Pressure Sensing 94


4-5 Sound Sensing 95
4-6 Flowrate Sensing 99

4-7 Electromagnetic Sensing 101

4-8 Temperature Sensing 104

4-9 Light Sensing 106


4-9.1 The Photoresistive Sensor, 106
4-9.2 The Photoemi&sive Sensor, 108
4-9.3 The Photovoltaic Sensor, 108

4-10 Other Control System Components:


Servomechanism Devices 108

4-11 The Servomotor 108


4-11.1 Calculating the Transfer Function
for a Servomotor, 111

4-12 The Gear Train 120

4-13 Rate Generators 122


4-13.1 The Rate Generator s Transfer
Function, 123

4-14 Synchromechanisms 123


4-14.1 The Transfer Function
for Synchros, 125
4-14.2 The Synchro Control Transformer, 126

4-15 Resolvers 127

4-16 Auxiliary Servo Circuits 127


4-16.1 The Modulator, 128
4-16.2 The Demodulator, 130
4-16.3 The Phase-Lead Network, 132
4-16.4 The Phase-Lag Network, 133
Contents xi

4-17 The Stepper Motor 134

Summary 137

Exercises 138

5: BLOCK DIAGRAM MATHEMATICS 141

5-1 Breaking Up Systems into Blocks 141

5-2 Reviewing Old Terms and Learning New


Ones 144

5-3 Cascading Blocks 147

5-4 Eliminating Loops 148

5-5 Moving Summing Points 149

5-6 Interchanging Blocks 151


5-7 Combined Block Reductions 151

Summary 156

Exercises 156

BODE DIAGRAMS 160

6-1 Introduction 160

6-2 Bode Diagram Construction 160

6-3 The Elements of Bode Diagrams 171

6-4 How Accurate are the Straight-line


Approximations? 175

6-5 Graphing Complex Functions 175

6-6 Determining a System’s Stability 178

Summary 184

Exercises 184

CLOSING THE LOOP 186

7-1 Closing the Loop 186

7-2 The Closed-loop Gain Plot 186

7-3 Using the Nichols Chart 195


xii Contents

7- 4 An Example 197

Summary 203

Exercises 203

8: TRANSIENT ANALYSIS 205

8- 1 Introduction 205

8-2 A System’s Response to a Step Input 205


8-3 Generating and Analyzing a Transient Response
Curve 210
8-3.1 The Natural Resonant Frequency
and the Damping Factor, 212
8-3.2 The Damped Frequency, 214
8-3.3 Settling Time, 219
8-3.4 Number of Oscillations Before
Settling, 220
8-3.5 Over-shoot, 221
8-3.6 Response Time, 222
8-3.7 Handling Values of Step Inputs, 227

8-4 An Example Problem 227


8-5 The Components of a Transient Waveform 230

8-6 The Root Locus Method 230


8-6.1 Euler’s Equation and the Complex
5-Plane, 231
8-6.2 The Characteristic Equation, 233
8-6.3 The Routh-Hurwitz Stability Test, 237
8-6.4 The Root Locus, 239

8- 7 Other Controller Types 252


8-7.1 The Proportional Controller, 252
8-7.2 The Integral Controller, 253
8-7.3 The Derivative Controller, 254

Summary 255
Exercises 256

9: DESIGNING A CONTROL SYSTEM 258

9- 1 Introduction 258

9-2 Example 1: A Robotic Cart Steering


System 258
Contents xiii

9-2.1 The System’s Dead band, 261


9-2.2 Selecting the Components and
Generating the Open-loop Transfer
Function, 262
9- 2.3 Generating the Closed-loop Data, 267

9-3 Example 2: A Compass-directed Auto-pilot


Control System for a Boat 275

9-4 Example 3: A Remote Valve Positioner 280


9-5 Example 4: An Oscillograph 289

9-6 Example 5: A Conditionally Stable System 293

9-7 Example 6: A Digital Controller for a Hydraulic


Press 295

9- 8 Example 7: Bang-bang Servo Systems 299

Summary 304

Exercises 305

10: PROCESS CONTROL SYSTEMS 307

10- 1 Introduction 307

10-2 Robotic Control System 307


10- 2.1 Problems with Using Digital Encoding,
312

10-3 Vision Systems 314


10-4 Camera Leveler System 317
10-5 Paper Processing System 321

10-6 Liquid Flowrate Supervision System 326

10-7 Bottle Filling Process 331

10-8 Telescope Tracking System 333


10-9 Die Casting Process Control System 336

10-10 A Proportional Pneumatic Controller 340

10-11 A Remote Control Antenna Rotator 343

Summary 345

Review Questions 345


xiv Contents

11: SYSTEM COMMUNICATIONS 347

11-1 The Purpose of System Communications 347

11-2 Analog Data Communication 348

11-3 Carrier Modulated Control Systems 350

11-4 Amplitude Modulated Control Systems 350

11-5 Frequency Modulated Control Systems 353


11-6 Phase Modulated Control Systems 356
11-7 Digital Data Communication 356

11-8 Pulse Modulated Control Systems 359


11-8.1 Pulse Amplitude Modulation
(PAM), 359
11-8.2 Pulse Width Modulation (PWM), 359
11-8.3 Pulse Position Modulation (PPM), 360
11-8.4 Pulse Code Modulation (PCM), 361

11-9 Frequency Shift Keying 363

11- 10 Types of Transmission Media 365


11-10.1 Wire Conductors, 365
11-10.2 Radio Waves, 366
11-10.3 Fiber Optics, 367
11-10.4 Light Waves, 367
11-10.5 Pneumatic Transmission Lines, 367

Summary 368

Review Questions 368

12: COMPUTERS AND AUTOMATIC CONTROL SYSTEMS 370

12- 1 Introduction 370

12-2 What Is a Microprocessor 370


12-3 The Computer and Its Software 375

12-4 An Example of Automatic Controlling Using


a Computer 379
12-5 Using the Stepper Motor in an Automatic
Control System 387
12-6 Using a Programmable Controller for Automatic
Control 389
Contents w

12-7 Methods of Computer Interfacing 394


12-7.1 The 20 mA and 60 mA Current
Loops, 395
12-7.2 The RS-232C Standard, 396
12-7.3 The Centronics System, 399
12-7.4 The IEEE-488 General Purpose
Instrumentation Bus, 399

12-8 Plug-in Circuit Board Interfacing Methods 402


Summary 403

Review Questions 403

ANSWERS TO ODD-NUMBERED PROBLEMS 405

APPENDIX A: RECTANGULAR/POLAR
POLAR/RECTANGULAR CONVERSIONS
AND HOW TO HANDLE MATH
ROUTINES USING THESE FORMS 414

APPENDIX B: A COMPUTER PROGRAM


FOR PLOTTING TRANSIENT
WAVEFORMS 418

APPENDIX C: COMPUTER PROGRAMS FOR FINDING


CUBE ROOTS 421

APPENDIX D: EULERyS EQUATION 428

INDEX 431
Preface

This book is intended to be a textbook for students in engineering and engineering


technology. However, it is also intended to be a textbook for those engineers in
industry who find themselves involved in a project dealing with automatic controls
and would like to have a more-or-less self-taught refresher course. I can’t count
the number of times I found myself in exactly this position while in industry and
unable to find a practical book on the subject so that I could become an instant
expert and impress my peers. While I certainly don’t pretend to call myself an
expert on automatic controls, I hope that I have remedied this situation for a
substantial number of inquiring minds by writing this book.
I have always felt that physics was the great key to unlocking engineering
problems. This is one subject where theory is so important. I have my own
theory that says, if a person is observant enough, and if he or she can relate a
seemingly complicated technical explanation to an easily observed everyday liv¬
ing example, this would help in understanding that explanation and it would
certainly help in making that person feel more at home with the explanation. And
speaking of theories, the aforementioned is exactly the theory behind this book.
A fairly good background in basic physics coupled with a keen eye for observing
things will go a long way in understanding many of the concepts that I presented in
this book.
There is a fair quantity of math involved in my discussions of control the¬
ory. I have attempted to keep the math on a moderate level by circumventing
much of the differential and integral calculus normally used in dealing with this
subject. I feel that if one can learn a subject using a low-powered approach, why
destroy the interest in that subject by using a much higher powered approach? I

xvii
xviii Preface

do love calculus; it’s fun and illuminating to use. However, when it comes to
explaining everyday happenings, there are better approaches to take. Why use a
bulldozer in preparing a flower bed when a spade shovel will get the job done?
This book is intended to be a straightforward hands-on attack on automatic con¬
trols. I don’t believe calculus is the tool to use for that particular undertaking, at
least at this beginning stage. There are more important things to be understood at
this point.
Chapter 1 may appear to be an intense review of electronics, and that is
exactly what I intended it to be. I have attempted to cover as much electronic
theory as possible that deals with the area of controls. If you feel weak in this
area, this chapter will boost your electronics knowledge level to the point of
understanding the material that follows. You will undoubtedly find some portions
of this review not being directly applicable to the rest of the book. I believe,
however, that you can’t know too much about electronics when delving into
control systems. Someday you will need and want this additional information.
In Chapter 3 I have introduced a math concept for those of you who feel
especially comfortable with algebra. The concept is called Laplace transforma¬
tion. Laplace transforms are necessary in order to avoid differential and integral
calculus. Including this concept was certainly not an original idea of mine, since
other automatic control text authors have done it before me. However, I like to
think that I have applied these transformations to a greater extend than perhaps
some of the others have done. Using transforms is like using logarithms to avoid
multiplication and division. For those of you who are familiar with computers,
you might want to think of Laplace transforms as macros. Many of the more often
used calculus expressions have been transformed into these macros to reduce the
number of software statements and keystrokes. Transforms allow you to main¬
tain your sanity while working through some of the more difficult moments en¬
countered in automatic control system theory. They represent a very neat con¬
cept that is readily understood by most technical people. And a concept
understood is a concept that’s a friend for life, I always say.
Control systems are fun and impressive to watch. There is a certain amount
of beauty in watching a system perform a function and then make self-adjustments
and then continue functioning with no outside supervision. With the advent of the
microprocessor, the automatic control system is certainly approaching what is
best described as the closest thing we have for artificial intelligence (or real igno¬
rance, depending on its application). Chapter 12 deals specifically with this sub¬
ject of computers and automatic control systems.
In Chapter 4 you will find considerable material on electrical components,
namely, transducers. There is also an extensive discussion of the transfer func¬
tion. That discussion, along with the discussions on electrical transducers, is a
keystone to understanding the later material.
In the following chapters, I introduce many design problems along with their
solutions. I tried presenting them in a manner which I felt closely resembles the
thought paths taken by a typical design engineer. That is, the first idea for a
Preface xix

solution may not be the final solution used, but instead forms the basis for another
approach.
At the end of this book are appendices that contain some rather interesting,
but simple, software programs. If you have a computer, try them. You’ll have a
lot of fun with them, proving out the concepts discussed in the text.
I want to give special thanks to those individuals who have made this entire
writing venture possible. I thank professors Walter Anderson and Robert Stebler
for their vocal encouragement and for making much-needed writing resources and
secretarial help available. I also thank students Heather Baab and Todd Coulter
for the dedication and eagerness they showed in generating the much-needed
artwork this book required, using the School of Technology’s CAD facilities. It’s
young people like these that make engineering technology and teaching so enjoy¬
able and full of hope.
I sincerely thank one of my manuscript reviewers, P. Erik Liimatta, profes¬
sor of engineering and technologies at Anne Arundel Community College in
Arnold, Maryland, for his excellent comments and suggestions in reviewing this
manuscript. I know it was a time-consuming task, and I am indebted to his pa¬
tience which certainly contributed in making this a better book.
But most of all, I thank the members of my family. They showed tremen¬
dous patience while my many hours at the keyboard resulted in my being an
absentee husband and father. Certainly, without their encouragement, patience,
and understanding, none of this would have been possible.
And finally, it is my hope that you, the reader, will get as much enjoyment
out of reading and understanding this material as I had in writing it.

J.R.C.
Houghton, Michigan
AUTOMATIC
CONTROL
SYSTEMS
AND COMPONENTS
Review of Electronic
Fundamentals

1-1 INTRODUCTION

This chapter is meant to be a review of electronic circuit theory. It’s assumed that
you have a basic knowledge of atomic structure from a previous encounter with a
physics course or electronics course. This chapter reviews those basic principles
usually encountered in a typical automatic control circuit and presents some old
concepts in a somewhat different light for easier understanding. Few derivatives
are presented. These can be obtained, instead, by referring to any good basic
electronics text. Only the major highlights of circuit fundamentals are presented
in this chapter.

1-2 THE DC CIRCUIT

For all the debates over whether or not the use of water flow in a pipe to explain
the flow of electrical current in a wire is scientifically proper or not, the anology
does make the point. It is used here to demonstrate the many characteristics of
electron flow behavior. However, to begin this discussion, don’t think of a pipe
carrying water, but instead, think of a pipe carrying marbles. Assume that this
, pipe is completely filled with them. Now, if you were to place one more marble in
the pipe’s end, certainly, a marble will pop out of the other end, as seen in Figure
1-1. Note that this would not be the same marble that was placed in the pipe’s
other end. Current flow can be described similarly. A current flow is comprised
of electrons bumping or displacing adjacent electrons along a confined path such

1
2 Chapter l Review of Electronic Fundamentals

Figure 1-1 An electron flow analogy.

that a flow takes place inside this path. What causes this bumping or flow to take
place is discussed shortly. For the present, however, we use this analogy to help
explain current flow.

1-2.1 The Ampere

The electric charge found on an electron is the quantity that accounts for the
attractive and repulsive forces that exist between adjacent electrons and protons.
The electric charge carried by just one electron is 1.6 x 10"19 coulombs. It takes
6.242 x 1018 electrons multiplied by 1.6 x 10~19 coulombs to form one coulomb of
electric charge. This means that if 6.242 x 1018 electrons flow past a particular
point in a conductor in one second (that is, one coulomb each second), one
ampere of charge-flow is flowing. This flow, in amperes, can be found by the
following equation:

I Q
= (l-l)
t

where / = current (amperes, or simply “amps”)


Q = charge (coulombs)
t — time (seconds)

EXAMPLE 1-1 -----


Calculate the amount of current, in amperes, resulting from 39 coulombs flowing
through a conductor past a given point in that conductor, in 16 seconds.
Solution:
Using Eq. (1-1), the amount of current, in amperes, would be:
r/ . 39 coulombs
/(amps) = ——-
16 secs
= 2.438 amps

1-2.2 The Volt

In order to force electrons to flow through a substance, a pressure difference is


needed, just as a pressure differential is needed to force water through a conduit
such as a pipe. This pressure can be created through an electrochemical process
Section 1-2 The DC Circuit 3

(batteries) or it can be created electronically (electrical power supplies). In both


cases, a reservoir of electrons is created opposite a similar reservoir that has a
deficiency of electrons. Since this situation represents an unnatural or unstable
state (every chemically stable thing that occurs in nature is electrically balanced
or neutral, so to speak), the overabundant numbers of electrons will strive to find
a path back to the deficient electron area in an attempt to “balance things up”
once again. The pressure, or potential difference as it is sometimes called, needed
to separate these electrons represents an expended energy or work. If the unit of
energy is the joule, and the unit of pressure is called the volt, then the following
equation shows the relationship between the work expended, the coulombs,
moved, and the pressure needed to move these electrons:

0-2)

where V = potential difference or pressure (volts)


W = work expended (joules)
Q = charge (coulombs)

EXAMPLE 1-2
Assume that it required 136 joules of work in order to keep a charge of 12 coulombs
from recombining with the parent ionized atoms in a battery. Calculate the voltage
developed during this process.

Solution:
Using Eq. (1-2), calculate the voltage:

136 joules
V(volts) --
12 coulombs
11.333 volts

1-2.3 The Voltage Source

The electronic symbol used for a direct current voltage source (that is, a voltage
source that “pumps” or moves electrons through a circuit in one direction only) is
illustrated in Figure 1-2. The longer side of the symbol is understood to be the
positive (+) terminal, while the shorter side of the symbol is the negative (-)
terminal side. Remember, the negative side contains an overabundance of elec¬
trons, whereas the positive side has a deficiency of electrons. The voltage sup¬
plied by this source is usually stated alongside the symbol. It is important to note
that this source is usually considered ideal. That is, its voltage output never

+ i -
Figure 1-2 Electronic symbol for a volt¬
age source.
4 Chapter 1 Review of Electronic Fundamentals

varies. We think of battery voltages as “running down” after prolonged usage,


but not this voltage source. The reason for assuming a “perfect battery” for this
particular source is merely to simplify circuit performance characteristics. Things
can get fairly complicated if we work with a voltage source whose output varies
while we are trying to make measurements to prove a point.

1-2.4 Current Direction

It is important to pause for a moment to discuss a very important aspect of DC


circuit behavior. This concerns the direction of electron (or current) flow result¬
ing from a pressure difference (i.e., voltage difference) in a circuit. Remembering
that it is the electrons that do the traveling, so to speak, in a circuit, it’s logical to
assume that electron flow in any given circuit is from the negative side of the
voltage source to its positive side. This is logical, because the negative side
represents a packed accumulation of electrons waiting to break out when given
the chance. Of course, in order for these electrons to travel from the (-) side to
the (+) side, they have to go by way of any circuit devices that happen to be in
their path.
So what is current flow? Since many years ago it was thought that current
flow was from positive to negative (hence the expression, “shorting something to
ground,” where ground was, and still is today, thought to be a negative sump or
reservoir that electrons readily flow into), this is now referred to as conventional
current flow. This convention is used consistently throughout this text. So from
now on, instead of referring to electron flow as flowing from (-) to (+), we refer to
conventional current flow, or simply current, that flows from (+) to (—).

1-2.5 Determining Polarity

Once the direction of current flow has been established in a circuit, the polarity of
the voltage drop across a device resulting from that current is readily obtained.
Look at Figure 1-3. The circuit devices through which current is flowing are all
resistors. We would like to know the polarities of all the voltage drops occurring
across each resistor. In order to do this, simply remember the rule of thumb:

Current goes in positively and comes out negatively.

+
E I
t

Figure 1-3 Current flow through a


circuit of resistive devices.
Section 1-2 The DC Circuit 5

Using this rule, the polarities developed across each resistor would look like
Figure 1-4. This rule can also be used in reverse. If the polarities are given across
the circuit components and you had to determine the current directions, merely
determine the directions that would generate these polarities.

Figure 1-4 Resultant voltage polarities


of current flowing in a circuit.

1-2.6 Ohm’s Law

The relationship that exists between voltage and charge-flow, as the voltage (pres¬
sure) forces the charge down a conductor, depends on how much of an impedi¬
ment that the conductor’s material decides to create for the flow of current.
Obviously, no substance is going to allow a flow of electrons, especially if it’s a
large flow, to move through it without creating some sort of internal resistance in
the process. This resistance is in the form of friction, and friction generates heat.
(This problem is discussed in detail in Section 1-13.) Let’s now take a look at the
relationship that exists between this resistance, which has been given the unit of
the ohm, the volt, and electron flow or current. That relationship is expressed in
the following equation, called Ohm’s law:

'■i "-3|
where I = current (amps)
E = voltage (volt)
R = resistance to current or electron flow (ohm)

EXAMPLE 1-3
Find the current, in amps, flowing through a 270-ohm resistor that experiences a
voltage drop of 17.4 volts.

Solution:
> Using Eq. (1-3), the current would be equal to:

17.4 volts
/(amps) =
270 ohms
= 0.064 amps
6 Chapter 1 Review of Electronic Fundamentals

1-2.7 Resistivity

The electrical resistance of any material is determined by four major factors.


These are:

1. the type of material (i.e., its chemical make-up),


2. its temperature,
3. its length, and
4. its cross-sectional area.

All four of these factors are related to one another by the following expres¬
sion:

R = -r~ d-4)

where R — resistance (ohms)


k = resistivity factor or proportionality constant, which is temperature
dependent (ohms-circular mils/ft, or ohms-circular mils/meter)
Acm = cross-sectional area (circular mils, or CM. See following discussion.

Table 1-1 lists the resistivity constants, expressed in the English system of
measurement, for several common materials.
In order to convert any given area into units of circular mils (CM), the
following equation can be used (remember that a linear mil, or simply “mil,” is a
linear measurement in inches that has been transformed to thousandths of
inches. This is done by merely multiplying the linear amount by 1,000. The area,

TABLE 1-1 RESISTIVITY


CONSTANTS FOR SEVERAL
COMMON MATERIALS (VALUES
ARE FOR 68° F)1
Material Resistivity (ohms-CM/ft)

Silver 9.9
Copper 10.37
Gold 14.7
Aluminum 17.0
Tungsten 33.0
Nickel 47.0
Iron 74.0

1 Donald G. Fink, Electronics Engineer’s Handbook, (New York, N.Y.: McGraw-Hill. Inc.,
1975), pp. 6-4, 6-5.
Section 1-2 The DC Circuit 7

expressed in square mils, can then be found by multiplying the linear mil figures as
you would if you were using normal measurement figures):

(1-5)

where ACm — area expressed in circular mils (CM)


y4sq. mil = area expressed in linear square inches (in2) multiplied by 1,000,000
or 106

EXAMPLE 1-4 --
Find the resistance of a copper bar having a length of 14.5 feet and whose width and
height are 3.28 inches and 2.13 inches respectively.

Solution:
Step 1. Calculate the bar’s cross-sectional area using thousandths of inch units or
mils to calculate sq. mils:

Area = (3.28 in x 1,000)(2.13 in x 1,000)


- 6.986 x 106 mils2

(Notice that this is nothing more than taking the “normal” or linear area of 3.28 in. x
2.13 in. and multiplying the result by 1,000,000 or 106.)

Step 2. Convert the linear sq. mil area to CM area:


Using Eq. (1-5):

Acm = (6.986 in2) U x 106)

= 8.895 x 106 CM

Step 3. Determine the resistivity of copper:


Referring to Table 1-1, the resistivity constant for copper at 68°F is 10.37 CM-
ohms/ft.

Step 4. Calculate the resistance:


Using Eq. (1-6):

(10.37 CM-ohms/ft)(14.5 ft)


R ~ 8.895 x 106 CM
- 16.9 x 10~6 ohms
or 0.0000169 ohms

Once the concepts of current, voltage, resistance, and resistivity are under¬
stood, the next concept to review and understand is the DC circuit. Refer to
Figure 1-5. This figure shows several devices that utilize varying amounts of
voltage and current for their proper operation. One such device, the resistor, is
now described.
8 Chapter 1 Review of Electronic Fundamentals

*3

*4

Figure 1-5 A typical DC circuit.

1-3 THE RESISTOR

The resistor is a device specifically designed to be inserted into a conductor to


impede the flow of electrons in that conductor. The resistor can be constructed
either from a length of conductor itself, or it can be manufactured from a wide
variety of material mixtures whose electrical conductivities can be tightly con¬
trolled by varying the mixture proportions. A common mixture is comprised of
carbon-like material that is highly conductive, with a clay or ceramic-like sub¬
stance that is highly resistive to current flow. Varying the mixing proportions of
these two substances will produce varying amounts of resistances. The resistive
value associated with each resistor can be ascertained from a series of colored
bars printed directly on the body of the resistor if the resistance value itself isn’t
already printed there. Usually there are four bands. The first three bands indicate
the resistance, while the fourth band indicates the manufacturing tolerance. The
first two bands are the significant digits, while the third band is the decade multi¬
plier. Table 1-2 shows the color coding scheme that is widely used.

TABLE 1-2 THE


RESISTOR COLOR CODE

Black 0 Green 5
Brown 1 Blue 6
Red 2 Violet 7
Orange 3 Gray 8
Yellow 4 White 9
Gold 0.1
Silver 0.01

When the gold and silver bands are the third band on the resistor, they act as
multipliers, as shown previously.
Section 1-4 Series and Parallel Resistances and How They Are Calculated 9

Tolerances: Silver ± 10%


Gold ± 5%

When the gold and silver bands are the fourth band on the resistor, they act as
tolerance indicators, as shown previously.

EXAMPLE 1-5 -
Find the resistance of a resistor having the following color bands: first band = red,
second band = red, third band = orange, fourth band = gold.

Solution:
Referring to Table 1-2, the resistor would have a resistance of 22,000 ohms (first two
digits are 2 and 2, and the decade multiplier is 3 or 103, which is 10,000). The
tolerance of this resistor would be ±5%.

The physical size of a resistor has a considerable bearing on the amount of


current and voltage it can safely handle before overheating and destroying itself.
We discuss this problem further in Section 1-14. Some commonly used sizes or
“wattages” are shown in Figure 1-6.

Figure 1-6 Resistor sizes used in


electronic circuits.

1-4 SERIES AND PARALLEL RESISTANCES AND HOW THEY


ARE CALCULATED

There are two major types of resistive circuits: the series circuit and the parallel
circuit. There is also a third kind of circuit which is a hybrid of the first two, a
combination of the series and parallel circuit.
10 Chapter l Review of Electronic Fundamentals

1-4.1 The Series Resistance Circuit

Resistors placed in series have the following characteristics: The total current (IT)
flowing in the circuit shown in Figure 1-7 flows through all the resistors, R\, R2,
R3, and R4. If you were to measure the current at point c, point d, or point e, you
would find the measured current to be the same as IT. This is also the same current
flowing back up into the negative side of the voltage source, Es. If Es were

*3

ft*

Figure 1-7 A series DC resistive circuit.

thought of as being a water pump, the water leaving the pump must be the same
water entering the pump, assuming no losses (leaks) along the circuit of course.
Furthermore, as IT flows through each resistor, a voltage drop occurs across each
resistor (Figure 1-8).

/
/

Figure 1-8 Voltage drops developed in


a series circuit.

Each voltage drop amount is directly dependent on me value of each R


across which the drop occurs. And furthermore, the sum of V) + V2 + V3 + V4
must equal Es. The preceding conditions are true only because the current enter-
Section 1-4 Series and Parallel Resistances and How They Are Calculated 11

ing and leaving each resistor is the same current amount. The condition just
described defines the series resistance circuit. Summarizing:

Es = V, 4- V2 + V3 + • • • V, (1-6)

also

Rj — R\ + R2 + R3 + ■ ■ • Rx (1-7)
where RT = total resistance in a series circuit as measured between termi¬
nals a and b in Figure 1-7 (ohms)
R\, R2, R3 etc. = individual series resistors (ohms)

EXAMPLE 1-6 --
Refer to Figure 1-9. Here we see five resistors all wired in series with each other.
Determine the total circuit resistance (being aware that if this were an actual circuit,
you would want to remember to remove the voltage source, Es, while this is being
done. This is so the internal DC resistance of the supply will not be included in the
actual measuring of the circuit’s total DC resistance.). The circuit’s total resistance
would then be measured with an ohmmeter placed between terminals a and b.
However, we want to calculate the total resistance, not measure it. Also, in addition
to this, we want to calculate the individual voltage drops across each resistance.

/?t = 27 ft R2 = 22 ft

R3 = 120ft

£o = 13.5

/?4 = 470 ft

Figure 1-9 Circuit for Example 1-6.

Solution:
Step 1. Verify that this is in fact a series circuit. Do this by noting that the total
current (IT) must flow through all the resistors in order to get back to the other side of
the voltage source, Es. This means that each resistor will experience the same
' circuit going into and coming out of itself.

Step 2. Determine the circuit’s resistance as measured between a and b assuming


that Es has been temporarily removed for this determination: Using Eq. (1-7), let
R\ = 27 ohms, R2 ~ 22 ohms, R3 - 120 ohms, R4 = 470 ohms, and R5 - 82 ohms.
Therefore,
12 Chapter 1 Review of Electronic Fundamentals

RT = 27 ohms + 22 ohms + 120 ohms + 470 ohms + 82 ohms


Rt = 721 ohms

Step 3. Determine the circuit’s total current, IT:


Using Ohm’s law (Eq. 1-3) and the fact that the total supply voltage in the
circuit is 13.5 volts, then,

_ 13.5 volts
7 721 ohms
= 0.019 amps,

or 19 milliamps

Step 4. Determine now the individual voltage drops across each resistor.
Since each resistor has a particular resistive value with a known current flow¬
ing through it, we can again use Ohm’s law (Eq. 1-3) to determine the individual
voltage drops. To find the voltage drop across R\: Since I = E/R, then E = IR.
Note: The designation for voltage in Ohm’s law is E, and V in other expres¬
sions. The two letters are often used interchangeably to mean the same thing. E is
often used for voltage sources, whereas V is used for voltage drops occurring across
circuit resistances or impedances.

V\ = (0.019 amps)(27 ohms)


= 0.513 volts

To find the voltage drop across R2:

V2 = (.019 amps)(22 ohms)


= 0.418 volts

To find the voltage drop across R3 :

V3 = (.019 amps)(120 ohms)


= 2.280 volts

To find the voltage drop across R4 :

V4 = (.019 amps)(470 ohms)


= 8.930 volts

And finally, to find the voltage drop across R5:

V5 = (.019 amps)(82 ohms)


= 1.558 volts

According to Kirchhoff’s voltage law, all five voltages must equal the supply
voltage in this circuit. When totalling these voltages, we get 13.70 volts. We see that
this is not 13.5 volts, our supply voltage. The error resulted in our rounding off of
our five voltage calculations.
Section 1-4 Series and Parallel Resistances and How They Are Calculated 13

1-4.2 The Parallel Resistance Circuit

Refer to Figure 1-10. It would help to refer back to our water pipe analogy at this
point in order to envision the concept of parallel resistance. Es is now our water
pump supplying pressure to force the water through the various circuit devices (in
this case, resistors). Note that the total water flow (i.e., current IT) must divide at
point a, since there is a device (R\) allowing some of the current to return to the
intake side of our pump (the (-) terminal of Es) and a device, R2, that is similarly
situated allowing I2 to flow back to the pump, as does /3-4. Note that all three
current divisions recombine back at point b to again form IT before returning to the
pump. The point is this: If devices are connected across a voltage source such
that the (+) and (—) terminals are common to these devices, then these devices
are said to be in parallel with each other. Another observation is: If the voltage
source’s current divides and then recombines going to and coming from each
circuit device, then these devices are again said to be in parallel with each other.
Let’s take another look at Figure 1-10. Are resistors R2, R2, and R4 in
parallel with each other? The answer is no. R2 and R4 have the same current
flowing through them; consequently, R2 and R4 are in series with each other.
However, it is proper to say that R2 plus R4 are in parallel with R2 (and R\ too, as
far as that goes).

*3

Ft 4

Figure 1-10 A typical parallel resis¬


tance circuit.

To calculate the equivalent resistance of resistors that are in parallel with


each other, the following equation should be used (see Figure 1-11):

1111 1
— — — _l_ — -|- — • • • —
(1-8)
Rt Ra Rb Rc Rx

R.

_l Figure 1-11 Circuit used for Eq. (1-8).


14 Chapter 1 Review of Electronic Fundamentals

where RT = the total equivalent resistance that could replace all the paral¬
lel resistances without affecting the circuit (ohms)
Ra, Rb, Rc, etc. = the individual parallel resistors (ohms)

EXAMPLE 1-7 --
Refer to Figure 1-12. Calculate the total equivalent resistance that would be mea¬
sured across terminals a and b, and determine the voltage drops across each resistor.
Note: Remember to imagine Es removed from the circuit during this resistance
measuring so that its internal resistance doesn’t interfere with the measuring. You
would normally remove Es if you had to make this resistance measurement in an
actual circuit.

12fi

Figure 1-12 Circuit for Example 1-7.

Solution:
Letting the 4-ohm resistor = Ra, the 8-ohm resistor = Rb, and the 12-ohm resistor =
Rc, we can use Eq. (1-8) to solve for RT:

= 1 \ J_
Rr ~ 4 + 8 + 12
= 0.458

Rt = 2.183 ohms

Another equation can can be used for calculating the total parallel resistance of
two resistors, is the following:

RgRb
(1-9)
Ra + Rb

Equation (1-9) can be modified for larger numbers of parallel resistors; how¬
ever, the equation’s complexity increases to a point of unwieldiness. Consequently,
for circuits involving more than two parallel resistors, use Eq. (1-8) instead.
As for the voltage drops across each resistor, since these resistors are all in
parallel with each other, they will experience the same voltage drop. In other words,
the voltage drop will be the source voltage in this case, which is 24.2 volts.
Section 1-4 Series and Parallel Resistances and How They Are Calculated 15

1-4.3 The Combined Series-Parallel Resistance Circuit

This type of circuit is shown in Figure 1-13. This is where a combination of series
and parallel resistors are found, and perhaps more realistically describes a typi¬
cally encountered circuit of resistors. If the circuit in Figure 1-13 is understood,
then you shouldn’t have any further problems with understanding series and
parallel resistors and with being able to distinguish one circuit type from another.

/?, = ion

8fi

i2n

Figure 1-13 Circuit for Example 1-8.

EXAMPLE 1-8 --
Find the magnitude of current IT in Figure 1-13.

Solution:
In order to find IT, you must first find the total equivalent resistance, RT. Then,

IT = (Ohm’s law) (1-10)

But in order to find RT so that Eq. (1-10) can be solved, we begin by combining series
resistances and parallel resistances so that the circuit can be reduced to a simpler
form.

Let R] || R2 — Ra

Note: When you see the designation || used in conjunction with resistor discus¬
sions, the symbol means “in parallel with.”

, _ ^1^2
' Then
A Ri + R2

(10)(15)
therefore,
A 10 + 15
= 6 ohms
16 Chapter l Review of Electronic Fundamentals

Similarly, we let R4 | R5 = Rb

_ ^4^5
Then Rb
R4 + Rs

= (2Q)(2Q)
therefore, Rb
20 + 20
_ 400
40
= 10 ohms

Now let /?6 + = Rc (these resistors are in series).

Rc = 8 ohms + 12 ohms
= 20 ohms

The resulting simplified or reduced series-parallel circuit now looks like the
circuit shown in Figure 1-14. Upon inspecting Figure 1-14 we see that we can do

Ra = 6

Rc 20 n

Figure 1-14

some more circuit reduction. We can combine R4 with RB since they are in series.
Therefore,

let + Rb — Rd

or Rd = 6 ohms + 10 ohms
— 16 ohms

which now results in a more simplified circuit shown in Figure 1-15. Since RD is now
|| with Rc, we can combine those two to get

R A = 6 £1
—vwv-

25 V < Rd = 16 SI iRc = 20Sl

Figure 1-15
Section 1-5 Kirchhoff’s Laws 17

(16)(20)
Rf =
16 + 20
= 8.889 ohms

resulting in the further reduced circuit of Figure 1-16.

Ra - 6 f2

/?£■ — 8.889 £2

Figure 1-16

Since RA and RE are now in series with each other, we can combine those two
resistors to give us a final equivalent resistance of:

/?equiv = 6 ohms + 8.889 ohms


= 14.889 ohms

Essentially, we are saying that if all the resistors of Figure 1-13 were replaced
with a 14.889-ohm resistor, the voltage source, Es, would not know the difference
since it would experience the same current, IT, being drawn from it. However, we
still haven’t finished our problem here. We still have to find the magnitude of IT. By
going back and using Eq. (1-10) we can now calculate IT as

_ 14.889 ohms
T 25 volts
— 0.596 amps (or 596 mA)

1-5 KIRCHHOFF S LAWS

We now review two very important rules or laws developed by Gustav Kirchhoff
back in the mid-nineteenth century. These laws have to do with the behavior of
circuit voltages and currents.

1-5.1 Kirchhoff’s Voltage Law

Earlier it was stated that the sums of all the voltage drops occurring across
resistors in a circuit must be equal to the supply voltage at the voltage source.
This is because of Kirchhoff ’s voltage law. His law states that the algebraic sum
of the potential rises and drops around a closed loop in a circuit is zero. Let’s
take a look at Figure 1-17. We see three resistances, R\, R2, and R3 all in series
and wired to Es. The voltage drop polarities and values have all been determined
18 Chapter 1 Review of Electronic Fundamentals

v, = itr,
= (4) (10)
= 40 V

V2 = ItR2
= (4)(20)
140 V = 80 V

Figure 1-17 Using Kirchhoff’s voltage


law.

based on the direction and magnitude of IT, the total current. Note too that we are
working here with a closed loop of current flow. That is, the current, IT, loops
back to where it “began.” Selecting a point, say point a on the circuit as a
starting point, and picking a direction of summation, say counterclockwise as
shown, begin summing all the rises and drops in voltages. We find, therefore,
that:

-140v (a drop) + 20v (a rise) + 80v (a rise) + 40v (a rise) = 0

Picking any starting point and traveling in either direction, clockwise or


counterclockwise, will always produce the same results according to Kirchhoff’s
voltage law.

1-5.2 Kirchhoff’s Current Law

Kirchhoff has a comparable law to his voltage law that deals with current. Simply
stated, it is this: All current flowing into and out of a junction comprised of
conductors must equal zero. Or, stating it even more simply, what goes into
something must also come out of it.
Refer to Figure 1-18. Note point a. This is a junction because of the joint
formed by three wires or conductors. The current amounts flowing into and out of
the junction have all been determined. Using Kirchhoff’s current law and assum¬
ing that currents flowing into the junction are (+) and currents coming out are ( —)
(we could just as easily have picked the opposite signs and directions if we wanted
to and we would not have affected the outcome of the answer); the following
results would be obtained:

+ 10A (entering) - 6A (leaving) - 3A (leaving) - 1A (leaving) = 0

Now pick point b to see if Kirchhoff’s current law works there also.
Section 1-6 The Voltage Divider Rule 19

Ft i

Figure 1-18 Using Kirchhoff’s current


law.

1.6 THE VOLTAGE DIVIDER RULE

When working with resistive networks in circuits, a handy way to calculate volt¬
age drops across resistors is by using the voltage divider rule. Figure 1-19 shows a
typical situation where this rule can be applied. The voltage divider rule states
that:

R out
Tout = V,in (Ml)
R in

where Vout = the divider’s output voltage (volts)


Vin = the divider’s input voltage (volts)
tfout the total resistance across which Vout is found (ohms)
^in = the total resistance across which Vin is found (ohms)

In the preceding example, Rout = R51| (R3 + R4) and R[n = R51| (R + R4) + R + 3 1

R2. However, 2?out could be comprised of any type of resistor configuration, not
just the one shown in Figure 1-19. The same can be said of Rm also.

R2

Figure 1-19 Using the voltage divider


rule.
20 Chapter 1 Review of Electronic Fundamentals

EXAMPLE 1-9
Find the output voltage, Vout, for the circuit shown in Figure 1-20.

10 12

1 Out
/,

/
/

Figure 1-20 Circuit for Example 1-9.

Solution:
We see that Vm = 100 v. We must then find Rm. This value happens to be the
circuit’s total resistance across which the 100 volts is appearing. It is equal to:

(10 || 10) + 20 + 50 + (5 || 16) = 78.81 ohms

To find Vout, we must determine Rout, which equals (5 || 16) + 20, or 23.81 ohms.
Finally, according to Eq. (1-11),

'23.8F
V,out 100
-78.8L
= 30.21 volts

1.7 POWER SUPPLY REGULATION

Earlier we stated that the voltage sources used in our example problems and
circuit explanations were ideal sources. That is, they are unlike our car battery,
where, when we start the engine with the headlights on, the battery’s terminal
voltage drops, thus dimming the lights. Our ideal voltage source would not lower
its terminal voltage, but would instead remain constant. In the case of our car’s
battery, the starter’s increased current demand placed on the battery causes the
voltage to drop at its terminals. Batteries have an internal resistance that is inher¬
ent and unavoidable in their construction. A voltage drop develops across this
internal resistance just as if it were an external resistance like other circuit resis¬
tors. Since the sum of all the voltage drops across the series resistances must
equal the voltage source’s voltage (this is Kirchhoff ’s voltage law just explained in
Section 1-5.1), the greater the voltage drop across this internal resistance due to
Section 1-7 Power Supply Regulation 21

increased current flow, the less voltage there is available at the voltage source’s
terminals. Look at the illustrations in Figure 1-21. Notice that when a current of
8 amps is demanded from the battery, this creates a voltage drop of 4 volts across
the battery’s internal resistance, because

Lfl(iNT) = (fLOADX^INl)
= (8)(0.5)
= 4.0 volts

^ LOAD i.25 n

(a)

Figure 1-21 A typical battery circuit, (a) Before load is applied, (b) After
load is applied.

And according to Kirchhoff’s voltage law, this leaves only 10 volts to drop across
the load resistance (14V — 4V = 10V). This is the same voltage that appears
across the battery’s terminals at a and b. The battery’s regulation is not good
enough to keep its output voltage at a constant value because of its internal
resistance. Obviously, it would be nice if it were possible to manufacture batter¬
ies with little or no internal resistance.
Our ideal voltage source doesn’t have a regulation problem. We purposely
stated that the voltage stays constant regardless of current demand in order to
simplify our problem solving. Once the problem is solved under these ideal condi¬
tions, we can then analyze the same problem using a not-so-ideal voltage source
such as a battery or electronic power supply to see how the circuit actually
behaves. But this complicates the problem considerably, and we’ll avoid it for the
present.
If we were to plot the voltage regulation characteristics of our car battery
versus an ideal voltage source, we would get something that would look like
Figure 1-22. Note that, in the case of the ideal voltage source, as the current
demand increases the voltage output of the source remains constant. However, in
the case of the car battery, as more current is demanded its output becomes lower.
The percent of voltage regulation is calculated using the following equation:

V NL ~ VFL
% VR x 100 (M2)
Vfl
22 Chapter l Review of Electronic Fundamentals

Terminal Voltage

Figure 1-22 Load characteristics for


an ideal and actual voltage source.

where % VR = percent of voltage regulation (%)


Vnl = the no-load voltage at the terminals (volts)
Vfl — the full-load voltage at the terminals (volts)

EXAMPLE 1-10 -
Using the information in Figure 1-21, calculate the percent voltage regulation for the
battery shown.

Solution:
Using Eq. (1-12), the percent voltage regulation would be calculated as follows:

VNL = 14v

VFL = lOv

14v — lOv
Therefore, % VR =-—- x 100
lOv
= 40%

/THE AC CIRCUIT

Our next review subject has to do with alternating current or voltage. Up to this
point, we have been dealing with current flow in one direction only. We now
discuss the condition in which current makes direction reversals in the circuit
where it is flowing.

1-8.1 Generating an AC Signal

There are only two major methods used for generating an alternating current or
voltage. One is by using rotating mechanical machinery and the other is by using
electronic oscillators. Figure 1-23 shows a mechanical device, a generator, or
Section 1-8 The AC Circuit 23

alternator, in which a looped inductor is rotated through a magnetic field thereby


generating an electric current within the loop. The rotation direction of the wire
loop (F) and the magnetic field strength (B) determines the amount of current (/)
generated. The right-hand rule is an aid in remembering the relationship that
exists between the F, B, and I quantities. The revolutions per second made by the
rotating loop determine the frequency of current reversals within the circuit at¬
tached to the generator. If the loop is rotated at 3,600 rpm (revolutions per min¬
ute), the number of current reversals is 3,600 times per minute, or 60 times per
second. Each reversal is called a cycle, or hertz.
The second method of generating an AC signal is through the use of an
oscillator. There are hundreds of oscillator designs in use, and we won’t get into
that particular discussion, but again, any basic electronics design book is a good
source for these designs. In the oscillator, the frequency of oscillation is usually
determined by the value of a resistor, a capacitor, an inductor, or by an oscillating
piezoelectric crystal, or a combination of each. Each of these components is
reviewed later. For now, we want to review the properties and nomenclature
associated with AC waveforms.

1-8.2 Waveform Definitions and Conventions

The following is a list of definitions that apply to any alternating or repeatable


waveform. The sinusoidal waveform is used as a model for this discussion, but
any other periodic or repeatable waveform would do as well. (See Figure 1-24.)
24 Chapter 1 Review of Electronic Fundamentals

Figure 1-24 Important features of a cyclic waveform.

DEFINITIONS

WAVEFORM The trace or path made by a measurable quantity such as


voltage or current, plotted as a function of angular position, or time, pressure, or
temperature, etc.
INSTANTANEOUS VOLTAGE, CURRENT OR POWER The magnitude
of the quantity (i.e., the voltage, current, or power) measured at an instant in
time. This is synonymous with taking a snapshot of the waveform’s height at a
particular instant in time in order to record an otherwise fleeting value. Usually,
small letter symbols are used to represent these quantities, such as e, i, p for
instantaneous voltage, current, and power, respectively.
PERIODIC WAVEFORM A waveform that is cyclic; i.e., it repeats itself
after a period of time.
AMPLITUDE The maximum or peak value of a waveform. We use the
designation of Emax or 7max to signify this quantity in this book.
PERIOD The length of time between identical adjacent points of a repeat¬
ing waveform. The waveform must be traveling in the same direction for these
points to be identical. The capital letter T is used for this quantity.
CYCLE The portion of a waveform contained in one period of that wave¬
form. The unit of the Hz is used here.
FREQUENCY The number of cycles that occur in one second. Frequency
and period are related by the equation:
Section 1-8 The AC Circuit 25

f=j (M3)

Frequency has units of Hz, kHz, MHz, etc.

Because the electrical generator was generally recognized as the first of the
two methods used for frequency generation, it was common to plot waveforms of
current and voltage versus an angular function such as degrees or radians. The
angular function could then be related to the angular rotation per unit time of the
generator’s rotating armature or loop. This is still done to this day, not only for
mechanical generators but also for electronic oscillators.
The following mathematical relationships are often used in AC circuit anal¬
yses:

277 radians = 360° (See Figure 1-25) (1-14)

Arc Length = Radius

0 = 1 Radian (57.3°)

Figure 1-25 The radian.

1 radian = 57.3° (approximately) (1-15)


77 x degrees
To convert degrees to radians: radians = (1-16)
1 oU

degrees
Or, for a close approximation: radians = ^ (1-17)

, 180° x radians
To convert radians to degrees: degrees (1-18)
77

Or, for a close approximation: degrees = radians x 57.3 (1-19)


distance (rad)
Angular velocity, w, (rad/sec) (1-20)
time (secs)

277
Also, angular velocity, oj (1-21)

But since (a rearrangement of Eq. 1-13)


26 Chapter l Review of Electronic Fundamentals

where / = frequency in Hz,


then (o (rad/sec) = 2rrf (1-22)

The basic general mathematical format or “template” for a sinusoidal wave¬


form is:
y = Amaxsin(a)t) (1-23)

where y = the instantaneous amplitude on a waveform at cot


to = angular velocity (rad/sec)
t = instantaneous time (sec)
Amax = maximum amplitude of waveform (expressed in volts, current, power,
etc.)

EXAMPLE 1-11
Refer to Figure 1-24. Assume that the maximum amplitude of the displayed wave¬
form is 8.3 volts. The frequency is 35 kHz. Write the mathematical function that
describes the waveform having the given characteristics.

Solution:
Using Eq. (1-23) and Eq. (1-22), we write
y A max (277ft)
Since Amax = 8.3v and / = 35 kHz,
then y = 8.3jm(219, 912/)
or
y — 8.3sin(2.2 x 105/)

1-8.3 The Phase Angle

Quite often it is necessary to know how much a particular signal is “out of phase”
with another signal at a given time. As an example, consider the signal again in
Figure 1-24. Notice that it doesn’t appear to start at the zero axis of the v-t
coordinates, but instead appears to be shifted to the left somewhat. The amount
of shift is usually measured in radians or in degrees. If the shift is expressed in
radians, then multiples of tt are often used. (Remember, n radians = 180°; i.e.,
3.1416 rad. x 57.3 — 180 , approximately.) By convention, a sinusoidal wave¬
form shifted to the left is said to be leading the other waveform by so many
degrees or by so many radians. A sinusoidal waveform shifted to the right is said
to be lagging by so many degrees or by so many radians. Mathematically, Eq.
(1-23) can be modified to denote these phase shifts in the following ways:

For a leading waveform: y = AmdXsin(ajt + 0) (1-24)


Section 1-8 The AC Circuit 27

and for a lagging waveform y = Amaxsin(cot - 6) (1-25)

where 0 = phase angle shift (rad)

EXAMPLE 1-12
Refer to Figure 1-24 once again. Assume that the curve shown has been shifted 45
degrees (or 7t/4 radians) from the origin of the shown coordinates. Write the mathe¬
matical expression describing this curve. Assume a maximum amplitude of 4.3 amp
and a frequency of 60 Hz.

Solution:
Since the curve in Figure 1-24 has been shifted to the left to indicate a leading
waveform, we use Eq. (1-24). Also, we were told that Amax = 4.3 amps and 9 = tt!4
radians.

Then y = 4.3 sin (cot + 7774) (from Eq. 1-24)

and since / = 60 Hz
y
C4

and (from Eq. 1-22)


£
II

then w = 2tt60

- 377 rad

Therefore, y = 4.3sin(371t + 7t/4)

Note that the trigonometric function “cos” is frequently used instead of “sin.”
Since a cosine function waveform is 90° out of phase with a sine function waveform
(Figure 1-26), we could rewrite all Amaxsin(cot) functions by changing the sine to a
cosine and subtracting 90° or ir/l radians. In other words,

Amaxsinot = Amaxcos(cot — tt/2) (1-26)

Figure 1-26 The sin x and cos x curves.


28 Chapter l Review of Electronic Fundamentals

EXAMPLE 1-13
Rewrite the solution found in Example 1-11 so that the cosine function is used
instead of the sine function.

Solution:
Using Eq. (1-26), we rewrite the expression from Example 1-11, which was y -
8.3sm(2.2 x 105) to now read:

y — 8.3c6>s(2.2 x 105t — tt/2)

Now refer to Figure 1-26. Notice that shifting the cosine function to the right
will place that curve back in phase with the original sine function.

1-9 THE CAPACITOR

In the next few sections we discuss the characteristics of the capacitor, how it
functions, and how to calculate circuit total capacitance.

1-9.1 Capacitor Characteristics

The capacitor is a temporary electronic storage device that is synonymous with a


water storage tank for a municipality or building. The capacitor stores electric
charge. The unit of capacity is the farad, but because of it being such a large unit
of measure, we instead speak in terms of microfarads (abbrev. gE, or mF) and
micro-microfarads (abbrev. /x/xf, or mmf, or more recently, picofarads, pf). The
electronic symbol for a capacitor is shown in Figure 1-27. The earlier symbol is
still seen in some older schematics, while the later version is the one now being
used. The lower curved line of the later version’s symbol is usually connected to
earth or chassis ground when grounding becomes necessary. This curved side is
also the negative (-) side of the capacitor. Some capacitors are polarized and
require this polarized connection for their proper operation. Variable capacitors
are used primarily in electronic communications or in the controlling of timing
circuits or oscillator circuits in general.
The relationship between capacity, electric charge, and voltage, is the fol¬
lowing:

(1-27)

Earlier Version of Present Version of Variable Capacitor Figure 1-27 The various capacitor
Fixed Capacitor Fixed Capacitor symbols.
Section 1-9 The Capacitor 29

where C = capacity, farads (F)


Q = electric charge, coulombs (c)
V = voltage, volts (v)
Capacitance is also calculated using the following relationship:

8.85 x 1012 erA


(1-28)
C ~ d

where C = capacitance (F)


A = plate area in square meters (m2)
d = distance separating the opposite plates (meters)
er = relative permittivity, sometimes called the dielectric constant (no
units)
The dielectric constants of various substances used in the manufacturing of
capacitors are listed in Table 1-3.

TABLE 1-3 THE DIELECTRIC CONSTANTS


OF SEVERAL SUBSTANCES

Substance Dielectric constant (er)

Air 1.0006
Barium-strontium titanite 7500.0
Porcelain 6.0
Transformer oil 4.0
Bakelite 7.0
Rubber 3.0
Paper, parafinned 2.5
Teflon 2.0
Glass 7.5
Mica 5.0

EXAMPLE 1-14 -
Find the capacitance of two flat aluminum plates each having an area of 100 cm2 and
whose faces are parallel and directly opposite each other and separated by a sheet of
glass 0.15 cm thick.

Solution:
From Table 1-3, we see that the er for glass is 7.5. Since Eq. (1-28) requires that the
area of the plates be expressed in meters rather than in centimeters, we must first
convert the given area to meters. Since there are 104 cm2 in a square meter, then
there are 100/104 m2, or 0.01 m2 of capacitor plate. Also, we must convert the glass
thickness, in cm, to meters. Therefore, 0.15 cm is the same as 0.0015 m. Therefore,
using Eq. (1-28),
30 Chapter 1 Review of Electronic Fundamentals

(8.85 x 10~12)(7.5)(0.01m)
C ~ 0.0015m
= 44.25 x 10-11 farads

or = 442.5 pF

An important characteristic of a capacitor is that it blocks a flow of DC


current. This is because of its construction. The capacitor is comprised of two
conductors separated by a nonconducting barrier or dielectric. A DC current
can’t pass through such a construction. However, a capacitor will allow a portion
of an AC current to pass through it. We will have more to say on this subject later
in Section 1-13.

1-9.2 Capacitors in Parallel

Capacitors that are installed in parallel are identified as such just as resistors in
parallel with each other. However, that is where the similarity ends. To calculate
an equivalent capacitance for parallel capacitors, all the individual capacitors are
added together. That is,
CT= CA + CB + Cc + ■ ■ ‘ cx (1-29)
where CT = the total capacitance (farads, pi, pf, etc.)
CA, Cb, Cc, etc. = the individual || capacitors (farads, pi, pf, etc.)

EXAMPLE 1-15 --
Refer to Figure 1-28. Calculate the equivalent capacitance for capacitors C%, C9,
C10, and Cn that are all in parallel with one another.

Figure 1-28 Circuit for Example 1-15.


Solution:
Using Eq. (1-29) we calculate:
Ct = C8 + C9 + Cj0 + C„

CT = 8 pf + 6 pf + 3 pf + 5 pf
= 22 pf
Section 1-10 The Inductor 31

1-9.3 Capacitors in Series

Capacitors that are installed in the series configuration are identified as such,
again, just as resistors in series. But again, the comparison must stop here. To
calculate an equivalent capacitance for series capacitors, you must treat them as
resistors in parallel by using the following equation:

1111
— — — _l_ — _|_ — + • • • —
1
(1-30)
Cr CA CB Cc cx
where CT = the total equivalent capacitance (farads, fiF, pF, etc.)
CA, CB, Cc, etc. = the individual series capacitors (farads, juF, pF, etc.)

EXAMPLE 1-16 - ~ “
Calculate the equivalent capacitance for the series capacitors Cj, C2, and C3 in
Figure 1-28.

Solution:
Using Eq. (1-30):
_1_ = J_ J_ J_
cT~ c{ + c2 + c3
= 1 J_ J_
_ 5 + 10 + 15

1 6+3+2
CT ~ 30

= 15
~~ n
= 2.727 pF

1-9.4 Series-Parallel Capacitance

The treatment of combination series-parallel capacitor circuits is similar to that of


series-parallel resistors. (Refer to Section 1-4.3 to review this process.) Using the
results from Examples 1-15 and 1-16, and combining them with the remaining
capacitors of Figure 1-28, you should be able to calculate the total equivalent
circuit capacitance measured between terminals a and b. This value will be 27.895
pF. Verify this result before going on to new material.

1-10 THE INDUCTOR

The next component that we study and review is the inductor. As you will soon
learn, there are similar behavior characteristics as compared to the resistor and
capacitor.
32 Chapter l Review of Electronic Fundamentals

1-10.1 Inductor Characteristics

The inductor is another temporary electronic storage device like the capacitor,
except that it is usually thought of as having a shorter term storage capability,
relatively speaking. The storage takes place in the electromagnetic field that sur¬
rounds the inductor when a current flows through it. An inductor is comprised of
one or more insulated turns of conductive material such as copper wire wrapped
around a center core having a known permeability. The permeability of a material
is an indication of how easy or difficult it is for magnetic flux lines to become
established in the material. As a result, there is a similarity between permeability
and conductivity in electrical circuits.
The symbols for an inductor are shown in Figure 1-29. The variable inductor
is used primarily in electronic communications.

Air Core Air Core Iron Core Variable


Inductor Inductor Inductor Inductor Figure 1-29 Inductor symbols.

The unit of inductance is the henry. Typical variations of the henry are
millihenries and microhenries.
The behavior of the inductor when encountering AC- and DC-type signals is
exactly opposite of that of the capacitor. The inductor, being comprised of a
continuous inductor wound around a central core, will readily allow a DC current
to flow through it. On the other hand, it will make an attempt to resist any change
of direction of current as is typical in an AC current flow. As far as the AC current
is concerned, it experiences an impedance to its flow just as if someone had placed
a resistor in its path. As we review later, the amount of impedance experienced
by the AC signal is dependent on the signal’s frequency in addition to the amount
of inductance possessed by the inductor.
The calculating of the self-inductance or simply inductance of a coil can be
done using the following equation:

N2ptA
(1-31)
i

where L = the inductance in henries (H)


N = number of turns of conductor
M permeability of core material in Webers/Amp turns-meter (Wb/At-m)
A = area of core in meters2 (m2)
/ = length of core in meters (m)
Section 1-10 The Inductor 33

EXAMPLE 1-17
Calculate the inductance of the coil in Figure 1-30. The permeability of air is 12.57 x
10~7 Wb/At-m.

Figure 1-30 Inductor for Example 1-17.

Solution:
Applying Eq. (1-31) and using the information given in Figure 1-30, we get:

(2502)( 12.57 x 10_7)(A)


L ~ 0.06

Note: The units have been dropped in the preceding equation to simplify its
solution. We continue to do this from now on.
We now have to calculate A, the area of the inductor’s core:

7rd2

tt(0.0072)
Therefore,
4
- 3.849 x 10~5 m2

(2502)(12.57 x 10_7)(3.849 x 10~5)


and finally,
0.06
= 504 x 10-7 H
- 50.4 /xH

1-10.2 Inductors in Series

Inductors in series are easily recognized as behaving just like resistors as far as
becoming mathematically combined. That is, their total equivalent inductance is
calculated just like totalling series resistances:

Lt = La + Lb + Lc + • • • Lx (1-32)
where LT = the total equivalent inductance in henries, millihenries, or microhen¬
ries (H, mH, fiH)

No example is given here to demonstrate solving a series inductive circuit,


since the solution method is virtually identical to the method outlined in Section
1-4.1 for series resistances.
34 Chapter 1 Review of Electronic Fundamentals

1-10.3 Inductors in Parallel

Inductors in parallel with one another are recognized as being so, again, just like
resistors. Parallel inductors can be replaced with an equivalent inductor whose
inductive value can be calculated using the same form of equation as used with
parallel resistors. That is,

1111
-)- — _)_...
1
— — — — -\- —
(1-33)
Lt La Lb Lc Lx

where LT = the total equivalent inductance in henries, millihenries, or


microhenries (H, mH, H)
La, Lb, Lc, etc. = individual parallel inductors in henries, millihenries, or micro¬
henries (H, mH, /xH)

Again, no example is given here to demonstrate the solution of a parallel


inductive circuit, since the solution is virtually identical to the one used in Section
1-4.2 for parallel resistors.

1-10.4 Series-Parallel Inductors

The analysis of a combination series-parallel inductor circuit is identical to that of


a series-parallel resistive circuit. Consequently, no example is given here to
demonstrate a typical solution. Instead, refer back to Section 1-4.3 to see how
this is done, if this method is unclear to you.

1-11 THE TRANSFORMER

The next component we review is the transformer. This device is used for chang¬
ing AC voltage and current amounts in a circuit. The next three sections discuss
the transformer’s characteristics and its ratings.

1-11.1 Transformer Characteristics

Transformers are comprised of two inductors constructed close enough to each


other so that the flux field surrounding the one encompasses the windings of the
second inductor. In this manner, a voltage that has been developed across the
primary winding of a transformer can become inductively coupled to the second¬
ary winding (Figure 1-31). In order for the flux field to develop, however, we need
an alternating current or a changing current for inductive coupling to take place.
A direct current of constant magnitude will not create a changing flux field. For
this reason, transformers cannot transform DC voltages.
Note the symbol used in Figure 1-31 to denote an AC signal source. This is
the symbol that we use from now on to indicate a voltage source whose output is a
Section 1-11 The Transformer 35

Primary Secondary

Figure 1-31 Magnetic coupling of a


transformer’s coils.

varying AC signal, either in the form of a voltage output or a current output. You
will always be told what form of output is being generated whenever this symbol is
seen. You will see either an e or an i followed by its magnitude or value.
Transformers come in many sizes and shapes, as do all the other compo¬
nents that we have reviewed so far. However, transformers can be broken down
roughly into three groups. These are:

1. the iron-core transformer, used primarily for the purpose of transforming


significant quantities of current and voltage at audio or “line” frequencies;
2. the air-core transformer, used mainly for communication applications; and
3. The variable core transformer, which may be used for either power or com¬
munications applications.

1-11.2 The Turns Ratio

The following relationship holds true for a transformer’s primary and secondary
winding turns ratio and the voltages being induced:

(1-34)

where Ep the primary winding voltage (V)


Es the secondary winding voltage (V)
Np the number of turns of wire on the primary
Ns the number of turns of wire on the secondary
i

EXAMPLE 1-18
A transformer has 35 turns of wire on its primary winding. The winding is supplied
with 115 VAC. How many turns are necessary on the secondary to develop 40
VAC?
36 Chapter 1 Review of Electronic Fundamentals

Solution:
Using Eq. (1-34), and letting Np = 35 turns, Ep = 115 VAC, and Es = 40 VAC, then

115 35
40 " Np

Solving for Np:

_ (40X35)
p 115
= 12.2 turns

The following is the relationship existing between a transformer’s turns ratio


and the induced current:

Ir = N1 (1-35)
Is Np

where Ip = the primary winding current (A)


Is = the secondary winding current (A)
Np = the number of turns of wire on the primary
Ns = the number of turns of wire on the secondary

EXAMPLE 1-19 -
If the turns ratio (the number of turns on the primary divided by the number of turns
on the secondary) is 3 to 1, and the primary current is 12 amps, find the output
current.

Solution:
Using Eq. (1-35) and noting that Ip = 12 amps and Np/Ns = 3, then

12 = J_
Is ~ 3
Solving for Ns:

Is= 12 x 3
- 36 amps

1-11.3 Power Transformer Ratings

It is very important to remember that a transformer doesn’t “step-up” or “step-


down” power as it does to voltage and current. The power remains the same
regardless of what happens to the current or voltage.
In the rating of a power transformer, it is often the practice to mention the
term VA rating or volt-amp rating. This is the voltage and current values of either
Section 1-12 Transient Behavior 37

the primary or secondary windings multiplied together. In other words, the pri¬
mary VA rating must equal the secondary VA rating. However, the VA rating is
not the transformer’s true power rating. We review this fact in Section 1-14.

1-12 TRANSIENT BEHAVIOR

Both the capacitor and inductor, because of their signal storage capabilities, ex¬
hibit a rather unique charging or discharging characteristic (or filling or emptying
characteristic, if you prefer) which is momentary, called transient behavior. To
explain this, it’s better to start with the capacitor, since its behavior is a little
easier to visualize than the inductor’s transient behavior.

1-12.1 The R-C Circuit

Figure 1-32 represents, in simplified form, a type of circuit in which a capacitor, C,


is charged through a resistor, R, whose purpose is to slow down the charging rate
of the capacitor. If the resistor were missing and the capacitor were allowed to
charge directly from a voltage source, the capacitor would become instantly
charged. Assume that the switch, S, were thrown to position A. The charging
current, i, would have to flow through R to reach C. The higher the resistive value
of R, the slower the charging rate of C. Or, to word it another way, the slower
would be the voltage build-up, vc, of the capacitor. The time that it takes for vc to
build up to the source voltage, Vs, is

t = 5RC (1-36)

Time Constant (r)

Figure 1-32 The RC time constant.


38 Chapter l Review of Electronic Fundamentals

where t = time to fully charge capacitor (sec)


R = value of resistor (ohms)
C = value of capacitor (farads)

Note: Many textbooks on control theory and electronics state that t = 5RC, as
Eq. 1-36 shows. However, in theory, it takes at least six time constants to fully
charge a capacitor. In practice, though, it is commonly agreed that five time
constants is sufficiently accurate for most applications.
The time constant for the preceding circuit (sometimes referred to as an R-C
circuit), is simply:

r = RC (1-37)

where r = one time constant (sec)

At the end of one time constant, the capacitor has reached 63.2% of its fully
charged value. The charging curve takes on the appearance of curve 1 in Figure
1-32. The equation for this curve is:

y = 1 - e~th (1-38)

where r = R x C (sec)

If switch S in Figure 1-32 were thrown to position B, capacitor C would then


discharge at a rate already fixed by the same RC time constant. The discharge
voltage curve would look like curve 2 in Figure 1-32. The equation for this partic¬
ular curve is:

y = e~th (1-39)

where r = R x C (sec)

After one time constant, the capacitor will have lost 63.2% of its full voltage
and will be within 36.8% of being completely discharged.

1-12.2 The R-L Circuit

The discussion for the resistance-inductance circuit, or simply, R-L circuit, is


similar to that of the R-C circuit. Referring to Figure 1-32 once again, merely
change the capacitor to an inductor and place switch S back to position A. We
now monitor the voltage, VL, across the inductor, instead of Vc as we did with the
capacitor. When switch S is thrown to position A in Figure 1-32, the voltage, VL,
immediately becomes equal to the supply voltage, Es. The time it takes for the
flux field initially surrounding the inductor when the switch was closed to com¬
pletely collapse to zero, is

5L
t = (1-40)
R
Section 1-13 Reactance and Impedance 39

where t = time to fully collapse flux field, or, time for VL to drop from a value of
Es to zero (sec)
L = inductance (henries)
R = resistance (ohms)

The time constant for this L-R circuit is simply:

t = — (1-41)

where r = one time constant (sec)

As in the R-C circuit, at the end of one time constant the inductor’s voltage
has become reduced to 63.2% of its original full voltage, Es. The curve represent¬
ing this deteriorating voltage across the inductor resembles curve 2 in Figure
1-32. It’s important to note at this point that while VL was approaching zero, the
voltage across the resistor, R, was building up to a value equal to Es. Kirchhoff ’s
voltage law demands that this happen. We have a situation where, for all practical
purposes, the inductor is now acting as a short and only resistor R is present for
creating a voltage drop within our circuit.
Now let’s see what happens when switch S is thrown to position B. Since
VR = Es, and this continues to be true for an instant after switch S is in position 2
due to the current flowing through the inductor and the inductor current’s inability
to change instantly, the voltage across the inductor now becomes equal to the
voltage across R, except now it is opposite in polarity. However, the flux field
surrounding the inductor now begins to collapse. This reduction now produces a
decrease in VL, but reversed in polarity, across the inductor. The characteristic
of this decrease is similar again to curve 2 in Figure 1-32, except for the reversed
polarity just referred to. In other words, curve 2 must be “flipped over’’ so that it
is shown decreasing from a negative value up to zero as time goes on.

1-13 REACTANCE AND IMPEDANCE

Reactance and impedance are the two major factors that complicate AC circuit
analysis as compared to the relatively simple DC circuit analysis. Various compo¬
nents used in AC circuitry show a tendency to “react” to AC signals in varying
degrees depending on that component’s capacitive or inductive nature, and on the
frequency of the signal itself. This reaction is in the form of a blockage or
pseudoresistance. As a matter of fact, this pseudoresistance has been assigned
the unit of the ohm to emphasize the similarity. Unfortunately, this commonality
of units has also led to a lot of confusion for people just beginning in electronics.
This is where the similarity ends when comparing this so-called impedance of
current to the resistance of a resistor. To begin with, an ideal resistor does not
react to an AC signal other than presenting the same resistance to it as to a DC
40 Chapter 1 Review of Electronic Fundamentals

signal. Let’s now take a critical look at this apparent lack of reaction of a resistor
to an AC signal.

1-13.1 The Resistor’s Reactance to AC Waveforms

When an alternating current waveform is sent through a resistor, an alternating


voltage waveform develops across that resistor. Let’s assume that we have
placed an oscilloscope across the resistor to read this voltage drop, and another
oscilloscope to read current, placed in series with the resistor. (In order to do
this, you want to be sure to use the proper current shunts in conjunction with the
scope’s probes to create the voltage drop needed, since an oscilloscope is a
voltage, not a current sensing device.) Refer to Figure 1-33. The resultant wave¬
forms are drawn alongside the oscilloscope symbols depicting what we would
observe on that scope’s screen. As a matter of fact, we would see nothing un¬
usual. But we should point out that the two observed waveforms are in phase
with each other. In other words, as the current increases from zero to its maxi¬
mum value, so does the voltage waveform. This is an important point to remem¬
ber when we discuss the capacitor and inductor’s behavior.

1-13.2 The Capacitor’s Reactance to an AC Waveform

Now let’s substitute a capacitor for the resistor in our experiment by referring to
Figure 1-34. Notice the waveforms produced by the oscilloscopes. The voltage
and current waveforms are 90° out of phase with each other. While it may not be
evident in comparing the drawn waveforms, the voltage actually lags the current
waveform by 90°. The reason is surprisingly simple. When the current flow into
the capacitor is at a maximum, the voltage developed across the capacitor is at a
minimum. It’s only when the current is at minimum or zero does the capacitor

Wired as an
Ammeter

Figure 1-33 The current-voltage relationship for a resistor.


Section 1-13 Reactance and Impedance 41

Oscilloscope

Oscilloscope
Wired as
Ammeter

Figure 1-34 The current-voltage relationship for a capacitor.

have its fully developed voltage. This is because when the capacitor becomes
fully charged, a kind of resistance or back pressure is created such that the current
stops flowing. This is capacitive reactance. The capacitor’s size determines how
soon the capacitor fills, and the waveform’s frequency determines when the filling
rate is going to create this resistive reactance. So obviously, capacitive reactance
is dependent on at least two items that we have mentioned here: (a) The size of
capacitor used, and (b) the waveform’s frequency. The actual relationship is
expressed in the following equation:

*c = ^ (1-42,

where Xc = capacitive reactance (ohms)


f = frequency (Hz)
C = capacitance (farads)
~j = a 7-operator, which means that there is a 90° phase angle difference
between current and voltage. The minus sign indicates that capaci¬
tive reactance is drawn 90° downward in a phasor or impedance
diagram. (See Section 1-13.5.)

Whenever a reactance figure is stated for a component (in this case a capaci¬
tor) you must also be careful to state the phase angle. For a capacitor it is -90°.
For instance, a capacitor may have a voltage drop across it having the value of
5.34v Z_—90°.
' Quite often, Eq. (1-42) is written in the following form:

Since a) = 2trf (from Eq. 1-21)

then Z — 90° d-43)


Xc = Jc
42 Chapter 1 Review of Electronic Fundamentals

EXAMPLE 1-19 -
Assume that an AC current of 250 mA at a frequency of 400 Hz is passing through a
0.1 mf capacitor. Calculate the reactance produced by this capacitor.

Solution:
Using Eq. (1-42):
v — _ _ J_
c 2tt(400)(0.1 X 10"6)

Note that the 0.1 mf capacitor has been converted to farads according to the
requirements of Eq. (1-42).

Xc = —j3978.9 ohms

or, 3978.9 ohms A-90°

Remember that the reactance of 3978.9 ohms is not measurable with an


ordinary DC ohmmeter. This produced quantity, instead, is a reactance to an AC
current’s frequency for a given capacitance. If you were to measure the DC
resistance of this capacitor, or any capacitor, you would read an infinite resistance
on an ohmmeter (or close to it, since you might read a very high resistance due to
the capacitor’s dielectric material inside). A capacitor acts as an infinitely high
resistance to DC regardless of any current flow.

1-13.3 The Inductor’s Reactance to an AC Waveform

Let’s now substitute an inductor in place of the resistor in Figure 1-33. This is
shown in Figure 1-35. Again we see that a 90° phase shift has taken place between
the current and voltage waveforms. And again this is due to the “filling” of the
storage flux envelope surrounding the inductor. However, we have a radically

Oscilloscope
' Wired as a

Wired as
Ammeter

Figure 1-35 The current-voltage relationship for an inductor.


Section 1-13 Reactance and Impedance 43

different sequence of events going on here. In general, when a flux field is in the
process of expanding around the inductor, the lines of flux cut across the induc¬
tor’s windings causing an induced voltage of opposite polarity to develop. This
creates back pressure that tends to reduce or diminish any further current or
voltage build-up. When the flux field is at maximum build-up, that is, no more
windings are being cut by moving flux lines, current begins to flow again through
the inductor since the back pressure has now been relieved. The voltage mea¬
sured across the inductor is now zero (remember that this is a pure inductor with
no DC resistance associated with the coil windings). Now when the field col¬
lapses, the flux lines once again cut across the turns of inductor wire but now
generate a forward pressure that helps to pull or push the current through it.
Furthermore, this forward pressure or voltage represents a reversal in polarity
across the inductor, that is, a reversal as compared to the voltage’s polarity during
the flux field build-up. If all of this sounds complicated, just remember that we
now have a situation that shows the current lagging the voltage by 90°. The
amount of inductive reactance is dependent on the amount of inductance and the
waveform’s frequency. This relationship is expressed by:

= +J2irfL (1-44)
where XL = the inductor reactance (ohms)
/ = frequency of current being supplied (Hz)
L = the inductance (henries)
+j = a 7-operator, which means that there is a 90° phase angle difference
between current and voltage. The (+) sign indicates that the induc¬
tive reactance is drawn 90° upward in a phasor or impedance dia¬
gram.

EXAMPLE 1-20 -
Calculate the reactance created by a 20-mH inductor subjected to a current whose
frequency is 1.2 MHz.

Solution:
Using Eq. (1-44):
XL = +/2tt(1.2 x 106)(20 x 10~3)

Remember: All units must be in henries and Hertz.

XL = +7150.8 x 103 ohms

or, 150.8 x 103 ohms zL+90°

Often, Eq. (1-44) is seen written in the following form:


i

Since oj = 27rf (from Eq. 1-21)

then, XL - +jo)L

or, ojL A+90° 0-45)


44 Chapter 1 Review of Electronic Fundamentals

1-13.4 What Is Impedance?

Impedance is the term used to describe the resultant of combining inductive and
capacitive reactances with resistances. Look at Figure 1-36. Here we have a
combination of inductance, capacitance, and resistance all being supplied with a
10 VAC source at 490 Hz. The problem is this: How do you find the current i, and
how would you keep track of the phase shifts created by the combined effects of
the capacitor and inductor? We can’t neglect the effect of the resistor. Even
though we know that a resistor doesn’t cause a phase shift between the current
passing through it and the voltage dropping across it, it will definitely affect the
overall phase relationship between the output voltage and current of a circuit
containing both resistors, capacitors, and/or inductors. To solve this problem, we
first must devise a way of combining all three components into one component, so
to speak, much like we did in our discussion of DC circuits where we combined all
the resistors together and then, using Ohm’s law, found the total current. In the
case of an AC circuit, however, where inductors, capacitors, and resistors are all
involved, we combine them all into what is called the circuit’s total impedance.
And this impedance, which will have the unit of the ohm just as do reactance and
resistance, will have associated with it a phase angle just as we found in the cases
of the two reactances just studied. Assuming that we have found this total equiva¬
lent impedance at some phase angle, can we still use Ohm’s law to solve for the
current, /? The answer is yes. Ohm’s law will look like the following expression
for AC circuits:

E
I = 0-46)
z ze
where I = the current (amps)
E = the voltage (volts)
Z Z6 = the circuit’s impedance, in polar form (ohms)

Before we can proceed with our discussion, we must do some more review¬
ing of terminology concerning ./-operators, polar forms, and the difference be¬
tween i and I and e and E.

L = 0.1 mh ft = 120 ft
—C5W-—AAM—

e = 10 sin 3079f
OR C = 0.0006 juF
e - 10 V @ 490 Hz

Figure 1-36 An RLC circuit.


Section 1-13 Reactance and Impedance 45

1-13.5 Understanding the J-Operator

Reactance values and impedance values are considered vectors in electronics. A


capacitive reactance of, say —j 172, has both magnitude and direction, just as an
airplane flying at 225 mph NW has both magnitude and direction. Figure 1-37
shows the coordinate system for the ./-operator. Actually, most ./-operators have
associated with them a real number along with a ./-operator number, or what some
mathematicians call an imaginary number. A typical complete ./-operator plus real
number would look something like 110 — j 172. This entire number is often re¬
ferred to as a complex number. A plot of —j 172 and 110 — j 172 are shown in
Figure 1-37, which also shows their vector quantities. Notice the angle, 9. This is
the phase angle for 110 - 7172. The phase angle for simply —j 172 would be
-90°. To calculate the phase angle, 0, for any given complex number,

0 = ArcTan (1-47)
A

where ±A and ± jB are the components of any complex number.


To find the magnitude of any complex number (in our case, the magnitude of
the reactance of impedance),

Z = VA2 + B2 (1-48)

where, again, ±A and ±jB are the components of any complex number.
By now you may have guessed or recalled that the A term in our complex
number is a resistive value that always gets plotted on the R axis of our ./-operator
coordinate system. The B term is the reactive or impedance value that always

+/'

Figure 1-37 How complex and polar


forms are related.
46 Chapter 1 Review of Electronic Fundamentals

gets plotted on the j axis. Let’s now calculate both the magnitude and direction of
our example 110 - j 172. Using Eq. (1-47), we find the phase angle to be:

A ^ -172
6 = ArcTan ■ ■

= -57.4°
And by using Eq. (1-48), we can calculate the magnitude of the impedance:

Z = VllO2 + (—1722)
= 204.2 ohms
The usual convention used for writing our completed expression is: 204.2
Y-57.4°. This answer is called a polar form and is read “204.2 ohms at a phase
angle of minus 57.4 degrees.” If, at this point, you feel a little uncertain about
using j-operators, or phasors, as they are sometimes called, refer to Appendix A
at the end of this book to review the conversion process. The complex number
and 7-operator occur again in later chapters.

1-13.6 Instantaneous Versus RMS Voltages and Currents

Before we get back into our impedance discussion, there is one additional matter
of convention we must discuss and review. This concerns RMS and instantane¬
ous voltages and currents and how to distinguish one from the other.
Up to this point in our review of electronic circuitry, we have assumed that
whenever we were discussing alternating voltages or currents, we were discussing
the waveforms in their true size and shape as one would see them on, say, an
oscilloscope screen. If you were to hand-plot a waveform described by the ex¬
pression, y = 14sm(377f + 7t/4), you would graph a sinusoidal waveform whose
frequency was 60 Hz (/ = coI2tt) and whose maximum and minimum amplitude
was exactly ± 14 volts with a phase shift of 45°, leading. You would, in this case,
be plotting the instantaneous values of that particular waveform. Consequently,
whenever a waveform is presented in the form of y = Amsin(ajt ± 6) or y =
Amcos((ot ± 0), you will know that you will be dealing with instantaneous values.
As a matter of fact, notations using the small script e or i for voltage and current
respectively, are also an indication that instantaneous values are being used.
All of the preceding discussion is for the purpose of distinguishing instanta¬
neous voltage and current references from another kind of voltage and current
designation, called RMS voltage and current. RMS voltage and current (RMS
standing for the mathematical term root mean square) are used in discussions
dealing with power consumption or heat dissipation problems. This is where it is
necessary to determine an equivalent voltage or current amplitude that would
generate the same heat in a purely resistive network of resistors as would be
generated by a DC voltage or current source. As it turns out, for a sinusoidal
waveform, simply taking 0.707 times the maximum instantaneous value will give
you that equivalent heating value. This is called the RMS value, or sometimes the
Section 1-13 Reactance and Impedance 47

effective value. In the case of our preceding example where we looked at the
graphed instantaneous waveform of y = I4sin(377t + 7i/4), the RMS value would
be expressed as 0.707 x 14 or 9.898 volts Z+45°. Since we are speaking of RMS
values here, we used the polar form of the sinusoidal expression. That is the
convention we use in this book. So, whenever you see the polar form being used
to describe voltages or currents, you will know that we are speaking about RMS
or effective values.
The best way to understand all of these concepts is to consider another
example.

EXAMPLE 1-21
Solve for the current, /, in Figure 1-36.

Solution:
We solve this problem in a step-by-step manner.
Step 1. Find Xc: Using Eq. (1-42),

V z= _ _ 1
c 277(490 x 103)(0.0006 x 10~6)
~j
1.8472 x 10“3
= 0.541 x 10“3
- —j‘541 ohms

Step 2. Find XL: Using Eq. (1-44),

XL = 277(490 x 103)(0.10 x 10“3)


- 7*307.87 ohms

Step 3. Combine all the 7-terms and the real terms:

ZT = 7*307.87 - 7*541 + 120


= 120 - 7*233.13

Step 4. Find ZT by expressing in polar form: Using Eq. (1-48) and Eq. (1-47),

ZT = V1202 + (-233.32)
= 262.2 ohms

120
and 6 = ArcTan
-233.33
= zL-62.8°

Therefore, the impedance will be 262.2 Z-62.80

Step 5. Determine current, /: Using Eq. (1-46),

10
I =
262.2 Z.-62.80
= 0.038 zl+62.8°
48 Chapter 1 Review of Electronic Fundamentals

Notice that if we wanted to express our answer in an instantaneous value


instead, we would want to use the sinusoidal form, which would look like this:

i = 0.038 x 1.414ji/i(3.079 x 10bt + 62.8°)

or, i — 0.054sm(3.079 x 106/ + 62.8°)

where multiplying the RMS current by 1.414 converted it back to an instantaneous


value.
This resulting waveform is telling us that because of the impedance created by
the inductor, capacitor, and resistor in our circuit, the current is now being caused to
lead the voltage by 62.8°.
Impedances in circuit networks may be combined just as if they were resistors
in series or parallel using the same series and parallel equations as used for resis¬
tors. The only additional difficulty is the mathematical manipulating of the attached
phase angles with each impedance value. Again, it is suggested that you review the
information in Appendix A if you feel unsure of yourself in handling these kinds of
problems.

1-14 POWER

What makes a discussion on power so confusing is that there are so many different
types of power. The two power types that we are mostly concerned with are
the following.

DEFINITIONS

APPARENT POWER: The “power” resulting from merely multiplying


voltage times amps together as you would in a DC circuit. That is,

Pa = VI (1-49)
where Pa = apparent power volt-amps)
V = RMS value of the voltage (volts)
I = RMS value of the current (amps)

AVERAGE POWER, EFFECTIVE POWER, OR REAL POWER: This is


the power resulting from taking into account the phase angle existing between the
voltage and current waveforms.

P = pacos0 (1-50)
where P = average, effective, or real power (watts)
Pa = apparent power
0 = phase angle existing between the voltage and current waveforms (de¬
grees)
Section 1-15 Resonant Frequency 49

The term cos 6 in Eq. (1-50) is called the power factor. In general, electrical
equipment is rated in volt-amperes instead of watts, since it is not known by the
manufacturer what sort of electrical load or impedance is going to be attached to
the equipment by the customer. This would have a direct bearing on the phase
angle alteration that is most likely to happen when the equipment is installed. In
addition, it is quite possible for a customer to operate his or her equipment within
a seemingly safe power limit according to an attached wattmeter. However, the
current limitations placed on the equipment could still be greatly exceeded only
because the power factor was unusually low for some reason (perhaps due to a
highly inductive load). As a result, most manufacturers will rate their electrical
equipment in terms of volt-amperes (VA) instead of real power consumption,
leaving the real power consumption determinations up to the customer.

1-15 RESONANT FREQUENCY

The concept of resonant frequency plays a very important role in electronics and
in automatic controls in particular. A resonant frequency is one which is gener¬
ated by an electrical circuit (or mechanical circuit) that acquires a portion of its
operating energy through a more or less self-sustaining feedback loop or circuit.
Perhaps one of the best examples of such a system is shown in Figure 1-38. This is
a simplified diagram of an old radio circuit devised by an experimenter named
Armstrong back in 1922, called a regenerative receiver, or more appropriately, the
Armstrong oscillator. The theory of operation was quite simple and ingenious.
By merely “dumping” or feeding some of the amplified radio signal (amplified by
the rf or radio frequency amplifier) back to the antenna, the amplified signal could
become reamplified and reamplified again, until very weak signals could be de¬
tected and heard. However, if too much reamplification took place, the entire

Figure 1-38 The Armstrong regenerative radio receiver.


50 Chapter l Review of Electronic Fundamentals

circuit would oscillate or resonate at a particular frequency determined by the


radio’s tuning circuits. This was undesirable from the standpoint of radio recep¬
tion, but nevertheless an interesting side development for Armstrong. He later
used this principle for radio transmission experiments. We return to this example
later when we discuss automatic controls.
Any circuit that contains an inductance and a capacitance is usually a prime
candidate for resonating at a particular frequency. The principle is rather simple.
If a certain sized capacitor is allowed to discharge into an inductor, and the
inductor’s field is allowed to build up as a result and then to collapse so as to
reintroduce a charge-flow back into the capacitor, and then the capacitor is al¬
lowed to discharge back into the inductor, and so on, we have a resonant fre¬
quency condition. The frequency of this back-and-forth oscillating is determined
by the amount of capacitance and inductance in the circuit. The actual resonant
frequency can be calculated by the following equation:

(1-51)
^ IttVIC
where fr = the resonant frequency (Hz)
L = the inductance (henries)
C = the capacitance (farads)

EXAMPLE 1-22 -—
Refer to Figure 1-39. A signal generator wired as shown has its output frequency
varied over a range of frequencies. Determine the frequency that produces the high¬
est output voltage across the resistor as measured by the peak in the waveform seen
on the oscilloscope’s screen. In other words, find the circuit’s resonant frequency.

Figure 1-39 Determining the resonant frequency for an LC circuit.


Section 1-17 Solid State Devices 51

Solution:
Using Eq. (1-51),

fr =-
2ttV(50 x 10_6)(120 x 1012)
1
~~ 48.67 x 10~8
= 2.055 x 106 Hz

or, = 2.055 MHz

A frequency of 2.055 MHz from the signal generator will produce the highest
amplitude waveform on the oscilloscope’s screen.
Obviously, the back-and-forth oscillating of this electron charge cannot con¬
tinue forever by itself. Otherwise, we would have the makings of a rather contro¬
versial device, a perpetual motion machine. These have yet to be proven work¬
able. Internal circuit friction (resistance in this case), which is the usual downfall
of all such schemes, eventually puts an end to the oscillating action. However,
the oscillations can be continued for an indefinite period of time by simply inject¬
ing a fresh charge of electrons at just the right time from an outside source. This is
exactly what an oscillator does.

1-16 OSCILLATORS

The oscillator has valuable circuit applications in electronics, especially in the


area of timing. Oscillators become the circuit clocks that synchronize or coordi¬
nate the intricate functions of computer operations and of control circuits in
general. These types of precision clocks are usually further aided in their accu¬
racy through the utilization of crystals. These are precisely machined or etched
pieces of quartz-like material that vibrate at a resonant frequency either with or
without the aid of a resonant L-C circuit.
Another type of electronic oscillator that is frequently used takes advantage
of the R-C or R-L time constant. By varying the resistance of this type of circuit,
the circuit’s time constant can be varied over a wide range of values to simulate a
resonant frequency condition. Again, this can be used to produce the necessary
timing frequencies in many control circuitry applications.

1-17 SOLID STATE DEVICES

When working with automatic control devices, one can’t help but come in contact
with some sort of electronic solid state device being used in the control circuitry.
For the purposes of understanding automatic control theory, it isn’t necessary to
52 Chapter 1 Review of Electronic Fundamentals

fully understand the internal workings of these devices, although admittedly, it


helps. What we do here instead is deal with the devices pragmatically. Just
knowing in general how they work and what they do is certainly more than
sufficient for the purpose of obtaining a good grasp on how and why a system’s
design behaves the way it does.
Figure 1-40 shows the most often drawn symbols used in automatic control
designs, along with their identification. The following is a quick summary of what
each component does in a circuit.

DEFINITIONS

DIODE The diode is used for allowing current to flow in only one direc¬
tion, the direction indicated by the symbol’s arrow.

ZENER DIODE The zener diode is similar to the diode except that it is
used to regulate or maintain a particular voltage level in a circuit.

NPN AND PNP TRANSISTOR The transistor is used primarily for ampli¬
fying current signals. The transistor type (npn or pnp) determines the direction of
current flow within the emitter circuit as indicated by the emitter symbol.

FET TRANSISTOR This is a type of transistor that is a voltage signal


amplifier as opposed to the current signal amplifier qualities of the npn or pnp
transistor. The characteristics of a FET are similar to those of a vacuum tube.
There are several types of FET transistors; the type shown in Figure 1-40 is a
dual-gate MOSFET (the MOS in MOSFET standing for “metal oxide semicon-

Diode
+
Zener
Diode
NPN
Transistor
PNP
Transistor
Field-effect Transistor
(Dual Gate Mosfet)

NOR Gate NAND Gate Inverter

Integrated
Circuit

Figure 1-40 Typical solid state components encountered in automatic control


circuits.
References 53

ductor”). This device is used frequently in applications where it is necessary to


mix two signals to form a third signal. It is noted for its very small physical size.
NOR, NAND, INVERTER GATES These are logic devices designed to
make a logical decision based on a number of input choices given to them.

INTEGRATED CIRCUIT(IC) The IC covers a tremendously wide field of


devices too numerous to mention here. It is quite often difficult to determine how
a circuit that contains an IC behaves unless you are well versed in circuit design or
have used that particular device in another circuit. It is often necessary to refer to
reference manuals supplied by the IC’s manufacturer in order to find out how the
device functions. ICs are complete electronic circuits comprised of various com¬
ponents all packaged into a small epoxied container designed to be soldered into
or socketed into a larger circuit.

SUMMARY

Chapter 1 was intended to be a summary of basic electronic circuitry theory.


Enough theory was presented to enable you to develop more confidence in inter¬
preting the various automatic control system schemes that are presented later in
the coming chapters. We discussed the definitions of DC and AC current and
voltage, along with resistance, capacitance, inductance, reactance, and impe¬
dance. We also defined transients and their characteristics. We discussed trans¬
former action, power, and resonant frequency. We also identified the many dif¬
ferent solid state components and their behavior in an effort to help you
understand the electronic circuitry that we discuss later.

REFERENCES

American Radio Relay League, The ARRL Radio Amateur’s Handbook, Newing¬
ton, Conn.: ARRL Pub., 1987.
Boylestad, Robert L., Introductory Circuit Analysis, Columbus, Ohio: Charles
E. Merrill Publishing Co., 1977.
Jones, Thomas H., Electronic Components Handbook, Reston, Va.: Reston
Publishing Co., 1979.
The problem with learning many modern-day engineering concepts is that the
concepts can be quite intimidating if you are not acquainted with the theory that
goes along with them. Rather than be intimidated, a little knowledge about the
subject goes a long way in clearing up mysteries and making you feel a little more
at home with the subject. Automatic control theory can be just one of those
threatening subjects if not approached properly. Therefore, with that bit of phi¬
losophy in mind, let’s now try to relate the basics of automatic control theory with
some everyday observations.
You may be surprised to find out that the ordinary stereo amplifier used in
playing recordings is an excellent source of automatic control theory basics. It is
necessary to consider the amplifier’s electronics at this point, but we’ll keep it
relatively simple for the present so as not to lose sight of what we’re trying to
accomplish here.

2-1 FREQUENCY RESPONSE

Many of us, at one time or another, have had the unpleasant experience of listen¬
ing to our favorite recording played on an inexpensive amplifier or player system
and have remarked, “Wow! That really sounds awful!’’ What we were doing was
really criticizing that system’s inability to faithfully play back the audio frequen¬
cies whose true sounds we have memorized. We can thank our ears and brain for
that. Our ears have an amazingly flat frequency response to a broad range of
frequencies from about 16 Hz to about 20,000 Hz (Figure 2-1). The stereo manu-

54
Section 2-1 Frequency Response 55

Figure 2-1 Frequency response curve for the normal human ear.

facturer tries to match this response characteristic as closely as possible, or as


close as his or her design and manufacturing budget will allow. If the manu¬
facturer doesn’t do this, our ears can usually detect the difference in sound
quality.
Let’s look at Figure 2-1 a little more closely. Notice that in this figure we
have plotted frequency versus decibels. Notice especially that the scale used in
making this response curve is logarithmic. This is because the human ear is
logarithmic in its response to sounds. Also, because of the extremely large span
of frequencies that the ear responds to (20,000 Hz-20 Hz, or a span of 19,980 Hz),
the logarithmic scale allows the plotting of this large span of data. Another inter¬
esting feature of the ear is the extremely flat frequency response throughout its
entire hearing range. That is to say, there’s very little detectable variation of
loudness within this flat range.
Ideally, an audio amplifier will possess this same output characteristic. Any
variation in the amplifier’s output will surely be detected by the ear. Figure 2-2
shows a frequency response curve of an audio amplifier. In reality, a typical
frequency curve would not be quite as smooth and symmetrical as the one shown
here. Notice that the curve shows amplification versus frequency. (Ignore the
second vertical scale alongside the graph for now. We talk about that one later.)
Amplification is often referred to as gain. This is nothing more than a ratio of the
amplifier’s output voltage divided by its input voltage. The question is, how much
gain variation can the amplifier get by with without the human ear being able to
detect it? The ear can just detect a variation of two or three decibels, but what
about gain? Equation (2-1) shows the relationship that exists between voltage
gain and decibel output:

Decibel output = 20 log (2-1)


L Fin J
56 Chapter 2 Control Theory Basics

Frequency (Hz)

Figure 2-2 Frequency response curve of a typical audio amplifier.

Consequently, in order to define frequency response, we must set up some


sort of criteria concerning the flatness of the response curve so that we know what
portions of the curve to include in our frequency response statement. In other
words, we have to know how much gain variation the ear can tolerate before it
detects noticeable and annoying changes in gain. Up to this point, our discussion
has concerned audio amplifiers. However, frequency response is a concern for
other kinds of systems also. Mechanical systems have a frequency response char¬
acteristic just as pronounced as electronic systems. These systems are also capa¬
ble of generating a frequency response curve and have similar tolerable response
limits just as the audio amplifiers. However, instead of the ear being the judge as
to what is tolerable and what is not, our general opinion based on observation and
experience dealing with vibration and motion is used. As we’ll soon learn in the
discussions ahead, the criteria used in analyzing automatic control systems is
virtually identical to that used for the amplifier.

2-2 GAIN

We have already defined the term gain in our discussion of frequency response.
But to formalize the definition in the shape of an equation, we can say the follow¬
ing:

„ . output voltage
Gam = --—— (2-2)
input voltage
Section 2-5 Phase 57

Since it is generally assumed that the ear can just detect a 3-decibel (dB)
change in voltage level or loudness, you have only to place 3 dB into Eq. (2-1) and
solve for V0ut/Fjn to find that 3 dB represents a voltage ratio change of 1.414 or
0.707, depending on whether the change was an increase (i.e., +3 dB) or a de¬
crease (-3 dB). More often than not, the voltage ratio change will often be a
decrease occurring at the extreme ends of the response curve. As a result, these
points are often referred to as the curve’s 3-dB down points. (Refer again to
Figure 2-2.)

2-3 BANDWIDTH

The total span, or bandwidth, of frequencies in our audio amplifier includes the
frequency range spanned by the 3-dB down points in the amplifier’s specifica¬
tions. In other words, this adds an additional frequency range to the flat portion of
the curve. The ear will most likely be unable to detect this change in gain within
the 3-dB down points, and as far as the manufacturer is concerned, it’s an addi¬
tional bonus for them in their advertising! Therefore, as a formal definition, the
bandwidth of a device is the span of frequencies that is included within the 3-dB
down points of that device’s frequency response curve.

2-4 LINEARITY

There are a number of things that determine the bandwidth of a device and they all
have to do with the circuit design and the types of components used in the
device’s design and construction. The device’s linearity is determined by compo¬
nents such as diodes, transistors, integrated circuits, vacuum tubes, capacitors,
inductors, and diodes. In general, the linearity of a device such as an amplifier
refers to that device’s ability to faithfully reproduce the input signal at its output.
This is done in such a manner so that the only variation in the output signal as
compared to the input signal is its magnitude.

2-5 PHASE

The phenomenon called phase is often a source of confusion. In order to better


understand this term, spend a moment studying Figure 2-3. This figure represents
a test setup that can be assembled by anyone wishing to explore the characteris¬
tics of an audio amplifier in order to better understand the terms defined and used
up to this point. We see our audio amplifier with its input not only attached to a
signal source, but it is also attached to a dual trace oscilloscope. This is a kind of
oscilloscope that allows you to look at two waveforms from two different signal
sources simultaneously. The amplifier’s output is also wired to the scope such
58 Chapter 2 Control Theory Basics

Dual Channel

Figure 2-3 Hook-up for analyzing the performance characteristics of an audio


amplifier.

that we can now view both the amplifier’s input and output waveforms at the same
time. The amplifier’s input is wired to the output of an audio sine wave signal
generator having an output frequency capability of 1 Hz to, say, 30 KHz. Fur¬
thermore, let’s assume that the output audio signal from the amplifier has been
attenuated such that its amplitude or voltage output is identical to the input
signal’s amplitude. We can make this adjustment to the output’s gain level some¬
where in the amplifier’s middle frequency range where we know the amplifier’s
gain is flat. For instance, if we know ahead of time that our amplifier has a flat
frequency response up to, say, 20 KHz, we would then adjust the output gain to
match the amplifier’s input at approximately 10 KHz. Let’s now analyze our first
results appearing on the oscilloscope’s screen as depicted in Figure 2-4.

Figure 2-4 Input and output wave¬


forms that are in phase with each
other.
Section 2-5 Phase 59

As we compare the input and output signals at our middle frequency value
where we adjusted the amplitudes to be purposely the same, we really don’t see
anything all that unusual. As a matter of fact, the signals are identical in every
respect, indicating that the amplifier is faithfully reproducing the signal being
supplied to it. Moreover, the positive and negative peaks of both curves are
occurring at precisely the same spots; that is, the two curves are in phase with
each other.
Let’s now change the frequency output of our signal generator. Let’s re¬
duce the frequency to some very low value such as 10 Hz and again compare the
two resultant curves on our oscilloscope. Figure 2-5 shows us the outcome.
Immediately, we notice two things that have changed. For one, the output ampli¬
tude has not diminished when compared to the input; for another, there now
appears to be a shift in phase between the two curves. More precisely, we note
that the output curve is leading the input curve by some measurable amount.
Once again, let’s change the signal generator’s frequency to a now much
higher frequency representing the upper end of our amplifier’s frequency response
curve, say, 18,000 Hz. Figure 2-6 represents the results obtained with our oscillo¬
scope. We now see that the output waveform is lagging the input waveform and
the output is again considerably less in amplitude as compared to the input wave¬
form. We can now list at least two observable things that happened in our experi¬
ment while changing the frequency of our signal generator:

1. The output signal diminished considerably at both the low end and high end
of the audio amplifier’s frequency response range.

Figure 2-5 The output waveform


leading the input waveform.
60 Chapter 2 Control Theory Basics

Figure 2-6 The output waveform


lagging the input waveform.

2. The phase relationship between the two curves became either leading or
lagging at the ends of the response curve.

The results of our test, assuming that we recorded output amplitudes and
phase variations throughout the amplifier’s entire frequency range, can be summa¬
rized in the two curves shown in Figure 2-7.

Figure 2-7 Frequency response curves for an audio amplifier not using negative
feedback.
Section 2-6 Feedback 61

Up to this point, we have only observed a phase shift phenomenon but have
not really explained what it is or what causes it. To begin with, phase shift is
caused by components whose outputs are frequency sensitive, or to say it another
way, are frequency dependent. Examples are capacitors, inductors, transistors,
certain integrated circuits, etc. In other words, any electrical component that has
a nonlinear range of operation will display this phase shift characteristic. Phase
shift occurs because of the operating of these components in their nonlinear
range. There are certainly other reasons that cause phase shift too. One common
cause is the different lengths of propagation paths over which the two signals must
travel. If one signal path is longer than the other, then the two signals will arrive
at different times and will be out of phase with each other. In the case of our
amplifier, however, this probably isn’t the case. What is interesting about our
phase shift problem, though, is that it can be substantially reduced if not entirely
corrected. This can be done through the use of negative feedback circuits.

2-6 FEEDBACK

We are all familiar with feedback. If a microphone is placed too close to the
speaker, the audio coming from the speaker is coupled back into the microphone
to be amplified once again by the amplifier’s system. This produces a louder
speaker output which, in turn, is fed back once again through the microphone, and
the procedure of amplification is repeated. The result is a very loud howling or
screeching noise indicating that the audio system is out of control and needs
adjustment. This kind of feedback is referred to as positive feedback. It is a type
of feedback where the output signal is fed back in phase with the input signal,
either on purpose or by accident, as the input signal enters the system or ampli¬
fier. Let’s analyze this system a little more closely to see if we can describe what
is going on in mathematical terms.
Figure 2-8 depicts our amplifier with no feedback. It is simply an amplifier
with an input and an output having an amplification or gain of A. Em is the input
signal voltage and E0 is the output signal voltage. This system is called an open-
loop system. In other words, no feedback is present. The gain of any amplifier is
simply A = EjEm. Now let’s add a feedback circuit to the amplifier (Figure 2-9).
Notice in Figure 2-9 that we have installed a feedback path by using a circuit that
connects the amplifier’s output back to its input through another circuit labeled
EC (feedback control). The feedback control is nothing more than a variable
resistor that controls the amount of output signal being dumped back into the

Figure 2-8
62 Chapter 2 Control Theory Basics

System

Amplifier
In
(A) fou, -7^/-

/ / Swi tch

Out
FC
Out
(B)

7^/ 7^

Figure 2-9 A positive feedback system setup.

amplifier’s input circuit. Let’s call the gain of this particular circuit B. What is
especially important to notice here are the indicated waveforms that you would
find if you placed an oscilloscope at the locations indicated. We have also added a
switch at the amplifier’s input. Notice that when the switch is closed, the output
signal from the amplifier will be fed back through FC in phase with the original
input signal coming from the signal generator, and both signals, added together,
are fed back to the amplifier’s input.
Consider what happens when the switch is opened. With the switch in its
opened position, we have an open-loop amplifier. That is, no feedback is present
within the system. This is similar to the situation in Figure 2-8. However, when
the switch is closed, the amplifier is introduced to a feedback circuit having the
following characteristics beginning at the input of FC: (a) The input voltage of
circuit FC is the same as E0; (b) The output voltage being supplied back to the
amplifier’s input is equal to E0 times B (the gain of the feedback control circuit), or
simply BE0; and (c)

E = Ein + BE0 (2-3)


From Eq. (2-2) we know that E0 = AE for any amplifier. Therefore, in place
of E in this equation, let’s substitute Eq. (2-3) to get:

E0 = A(Em + BE0) (2-4)

= AEm + ABE0 (2-5)


Now let’s arrange Eq. (2-5) so that it reads EjEm = /(A, E, B). (That is,
EJEm is a function of A, E, and B):

AEm = E0 - ABE0 (2-6)

E0{\ - AB)
^in (2-7)
A
Section 2-6 Feedback 63

Eo = A
or (2-8)
Em 1 - AB

It’s important to remember that this equation is for a system that uses
positive feedback for its operation.
Let’s now take a look at some what-if situations using Eq. (2-8):

1. What if the amplifier’s gain, A, times the feedback control (FC) gain, B,
is less than one?
Answer: The gain of the system is increased over and above what it would have
been with the open-loop system depending on the magnitude of A x B.

2. What if A x B is exactly ecpial to one?


Answer: The system’s gain would be infinitely high causing the amplifier to just
go into smooth oscillation provided that the AB product is maintained just at the
value of one. Theoretically, one could then remove the signal generator at this
point and have a perpetual motion device. However, like all good ideas, there are
catches! In this case, because of system losses due to less than 100% efficiencies
of component performances, it is necessary periodically to add additional make¬
up energy to the system to keep it oscillating smoothly. Or think of it this way: In
order to keep a child’s swing swinging at the same amplitude or height of swing,
you must occasionally add an additional push, otherwise the swinging would
eventually run down and cease.

3. What if A x B is greater than one?


Answer: The system’s gain calculation would result in a negative number, and as
a result, can’t be defined. The system’s behavior that would result from this
condition would be described as uncontrollable, wild, or unmanageable. Oscilla¬
tions at this point could result ultimately in damage or self-destruction to the
system.

Systems that employ positive feedback for their operation are extensively
used today, but not for automatic control applications, as we will find out later
on. Positive feedback has widespread use in electronic circuitry in the design of
oscillators. Years ago, a circuit that was very popular in radio manufacturing was
the regenerative radio receiver circuit that used a tickler coil circuit shown in
Figure 1-38 in Section 1-15. This system was a positive feedback system (its
operation was described in general in Chapter 1 but is worth repeating here). The
circuit worked like this: The very weak radio signal to be amplified entered the
grid of the vacuum tube by way of the antenna and step-up transformer coils a and
be The so-called grid-leak detector circuit aided the vacuum tube in detecting the
radio signal while the vacuum tube also amplified the signal. The signal coming
from the plate of the tube was then fed back to the antenna coil by means of the
tickler coil to allow the amplified signal to be reamplified all over again. The
amount of feedback was controlled by the radio’s listener by adjusting the variable
64 Chapter 2 Control Theory Basics

resistor, R. R was adjusted for maximum volume or just to the point before
allowing the circuit to break into oscillation (i.e., just before A x B = 1). If too
much signal was fed back to the antenna, the radio’s output at the headphones or
speaker would emit loud and uncontrollable howls and screeching noises, indicat¬
ing oscillations and too much feedback. The radio had amazing sensitivity to
weak signals because of this circuit and was used for many years before ultimately
being replaced by the modern-day superheterodyne receiver circuit. Even today,
some people still argue that the regenerative receiver, because of the positive
feedback circuit, still has very definite and beneficial applications in weak signal
communications.
Now that we have explored the behavior patterns of a system with positive
feedback, let’s take a look at our same system but with a negative feedback circuit
installed. What is interesting here is the modification necessary to convert
the positive feedback system to a negative feedback system. All that is needed
is to reverse the wire connections out of FC going to the input of the amplifier
(Figure 2-10). With this wiring revision, the signal leaving FC will now be intro¬
duced to Exn 180° out of phase with Em. In other words, when Em is increasing in
voltage, the output of FC is decreasing, and visa versa. Let’s now see how the
EJEm gain equation is calculated for our system with this reversed behavior
pattern.
Because of the fact that BE0 is now out of phase with Em by 180°,

E = Ein - BE0 (note the sign change) (2-9)

Again, from Eq. (2-2), we know that E0 = AE for any amplifier. Therefore,
in place of E in this equation, let’s substitute Eq. (2-9) to get:

System

Figure 2-10 A negative feedback system setup.


Section 2-6 Feedback 65

E0 = A{Em - BE0) (2-10)

= AEm - ABE0 (2-11)

Now let’s arrange Eq. (2-11) so that it reads EjEm = /(A, E, B) as we did in
Eq. (2-5):

AEm — E0 + ABE0 (2-12)

E0{ 1 + AB)
Em (2-13)
A

E0 A
(2-14)
E'm 1 + AB

Equation (2-14) is used for calculating the system’s gain employing negative
feedback. Notice that the only difference between Eq. (2-14) and Eq. (2-9) is the
sign in the denominator. Be sure not to confuse them.
Now that we have defined positive and negative feedback both verbally and
mathematically, let’s go back to our discussion of the audio amplifier in Section
2-5 concerning the use of feedback in controlling the phase angle shift between the
amplifier’s output and input signals. By using negative feedback, we can nullify or
cancel out any phase shift tendencies in our amplifier by introducing this feedback
at the proper instances. This is precisely what is done in hi-fi amplifiers to give
them the broad frequency responses necessary to faithfully reproduce the sounds
of a record, tape, or compact disk system. Figure 2-11 shows the results of our
redesigned amplifier now using negative feedback. Compare these curves with
those shown in Figure 2-7. A notable feature of Figure 2-11 is the two humps

Figure 2-11 Frequency response curves for an audio amplifier using negative
feedback.
66 Chapter 2 Control Theory Basics

occurring in the gain curve. Because of the negative feedback circuitry added to
the amplifier, it operates as negative feedback circuitry only when the out-of¬
phase relationship is not too severe between the amplifier’s input and output.
This is certainly the case for the majority of the flat response region of the gain
curve. However, near its ends, as we already discussed, the phase shifting be¬
comes much greater. It becomes so much greater in fact that this circuitry be¬
haves momentarily like a positive feedback system resulting in an actual gain
increase for these frequencies. This characteristic may be more fully understood
and better appreciated after studying Chapter 7.

2-7 THE BLOCK DIAGRAM

Because electronic and mechanical circuit systems can become rather compli¬
cated and difficult to draw, it is often necessary to use a shorthand method of
depicting these systems in illustrations. Blocks are used to represent the various
parts of the system and to simplify the system’s description. Figure 2-12 shows
the block diagram of a closed-loop system with all the necessary labeling needed
to identify the system’s operation. Notice that this type of drawing is easier to
draw and to understand than if a mechanical drawing of a water valve with its flow
valve being controlled by a float sensor were presented instead. And notice the
similarity between the block diagram shown in this figure and those shown in
Section 2-4. In general, block diagraming can be used to break down any control
system into the more basic components of a system similar to the one depicted in
Figure 2-12. Let’s now take a look at these basic elements. These are shown in
Figure 2-13.
We can define the functions of the individual blocks of Figure 2-13 in the
following statements:

1. Input components: These are the components of the overall system that
have to deal with the system’s input signal. More importantly, these are the
components that produce some sort of reference signal against which the

Water Water
Intake Output

Water Figure 2-12 Block diagram of a simple


_l
Level water level controller.
Section 2-8 The Summing Point 67

Figure 2-13 A generic block diagram for an automatic control system.

output may be compared in order to determine which way a correction must


be made to correct the system’s behavior.
2. Error determining components: These are the components that produce the
error signal as the result of comparing the input signal to that of the system’s
output signal thereby resulting in either a difference signal or an additive
signal that is then sent on to the control portion of the system. Don’t assume
that the error signal is an undesirable signal. On the contrary, the error
signal is an absolute necessary constituent of an automatic control system.
3. Control components: These are the components that modify the error signal
to put it into the desirable form for outputting. Typical control components
would be voltage, current, or power amplifiers, mechanical linkages that
control position, and control windings of servo devices.
4. Output components: These are the components that actually produce the
output of the system.
5. Feedback control components: These are the components that produce a
feedback signal that is of the same form as the input signal, but is 180° out of
phase, and is a scaled version of the output signal such that the feedback
signal may be combined with the input signal to produce the error signal.

2-8 THE SUMMING POINT

Refer back to Figure 2-13 for a moment. The block representing the error deter¬
mining components is often referred to as the summing point. This is where the
input signal and feedback output signals are added algebraically to produce the
error signal. Figure 2-13 may be redrawn as shown in Figure 2-14 where the error
determining components block has been replaced now with a summing point
symbol. This is simply a circle with an X drawn through it with the input and
output signal polarities all properly labeled. It can be generally assumed at this
point that all signals marked negative are 180° out of phase with the positive-
marked signals. Notice that the feedback signal entering the summing junction is
negative. This is to indicate that the feedback being used in this system is nega-
68 Chapter 2 Control Theory Basics

Figure 2-14 The summing point.

tive and not positive. Also, the input and output component blocks have all been
consolidated into one block called the forward components block. This name is
derived from the fact that the consolidated boxes all had signals flowing through
them in a forward input-to-output fashion. Quite often, you will see this form of
block diagramming used rather than the slightly more detailed form shown in
Figure 2-13. Instead of noting in each block the type of components involved, you
will often find the block’s system transfer function written.

2-9 THE TRANSFER FUNCTION

The transfer function of a system is nothing more than the gain of that system. In
other words, the transfer function is defined by the following equation:

^ r r the system’s output „ 1C,


Transfer function = —:-;—:- (2-15)
the system s input

Let’s take a look at some examples. Let’s determine the transfer function
for an AC motor. To do this we must first identify the motor’s input and output.
The input would, of course, be an AC voltage. The output would most likely be
measured in rpm. Therefore, its transfer function units would be rpm/AC volts.
For another example, let’s write the transfer function units for a water valve. The
input would be an angular displacement, degrees. This would represent the angle
through which the valve stem must be turned to control the desired water flow.
The output would be the water flow itself, measured in gallons per minute, or
gpm. Therefore, its transfer function units would be, gpm/deg.
In both cases, there most likely would be a numeric value associated with
the units of the transfer function. For instance, in the case of the water valve, a
45° turn of the valve stem may produce a flowrate of 9 gpm. Consequently, its
complete transfer function would be stated as 0.2 gpm/deg. We’ll come back to
this subject of transfer functions in a later section.

2-10 AN AUTOMATIC CONTROL SYSTEM

Up to this point, we have covered many of the basic concepts used in understand¬
ing automatic control systems. Let’s now try our hand at analyzing a system just
Section 2-10 An Automatic Control System 69

to make sure we have a good understanding of all the concepts covered so far.
This is probably the best way to uncover questions and to present additional ideas
that are covered later.
Assume that you and I are hot-air balloonists wanting to design a system that
maintains our balloon at a constant altitude regardless of air temperature change
or air draft conditions. We realize that our balloon’s lift, and to a great extent its
flight altitude, is determined by the difference in the balloon’s air bag temperature
and the temperature of the outside air. We need a system that will automatically
modulate or turn the propane burner off and on when needed to maintain the
proper lift or altitude. Obviously, there are other factors too that determine the
balloon’s altitude, such as wind direction and solar heating of the air bag, so we
need a system th^t will compensate for all of these conditions. Figure 2-15 shows
just such a system. The principle of operation is based on the aneroid barometer.
This is a device that is sensitive to changes in air pressure and therefore responds
to slight changes in altitude. If the aneroid mechanism senses a decrease in air
pressure, it interprets this change as an increase in altitude. If it senses an in¬
crease in pressure, the altitude must, therefore, be decreasing. To get our system
functioning properly, we must hook up the mechanism normally used to move the
needle that produces the barometric pressure reading on its scale to the gas
regulator on the propane burner. Assuming that we have done this, the system
will work like this: We must first adjust the altitude input control for the desired
altitude that we wish to maintain before taking off. Also, the aneroid barometer
must be adjusted for the existing barometric pressure reading so that at ground
level, the propane burner’s valve is opened initially to create the needed take-off
lift. As the balloon begins lifting off, the aneroid barometer responds to the in¬
crease in altitude by gradually shutting off the gas supply to the propane burner.
The balloon’s ascent is therefore slowed until it reaches the desired height. At
that time, the gas supply has been completely shut off.
Let’s analyze what would happen if a sudden updraft of air came along to
disturb our nice, level flight path. The aneroid would respond to the sudden
increase of the balloon’s altitude by first noting that the burner is already shut off.

Manual Altitude Propane Burner


Adjust Air Bag -■ ^ "Lift u
Input Control and Bag Exhaust

Aneroid
Controller
Manual Adjust
to Control
Response of
System

Figure 2-15 The automatic altitude controller for a hot air balloon.
70 Chapter 2 Control Theory Basics

Therefore, the aneroid would send a command signal to the gas bag to open its
venting valve to allow hot air to escape. The balloon would then respond by
losing some of its altitude. The aneroid, sensing this altitude loss and detecting
that this altitude is less than the desired altitudes as set by the input control, would
turn the gas burner on again to compensate for the balloon’s decrease in altitude.
Again, the aneroid, sensing an increase in altitude and sensing that the balloon’s
height is again greater than the desired preset altitude (but not as great as previ¬
ously, since a partial correction was already made), opens the vent valve to let
more hot air escape, and so on. Each correction would create a diminishing
amount of the automatic correction needed until finally the balloon would again be
at precisely the desired altitude.
Figure 2-16 is a diagram of the flight path of our balloon. After assuming
level flight, notice the two possible flight response paths in the diagram as a result
of an updraft . One is for our balloon if we had no automatic controls whereas the
other path is for our balloon with automatic controls. Can you imagine what our
flight path would look like if our automatic system had a much quicker response
time to the updraft of air current? Or, what would our flight path look like if our
automatic system responded very, very slowly? Figure 2-17 shows these two
possibilities. As you study this figure, ask yourself these questions: What would
happen to the response characteristics of our control system if we increased the

/
/
Without Automatic
Controls

/
/ Desired Path
or Altitude
/ \

TJ
3

/ With Automatic
/ Updraft Controls
/

Launch
Time
Point

Figure 2-16 A comparison of flight paths for a hot air balloon with and without
automatic altitude control.
Summary 71

weight of our balloon or its payload? Is there a possibility that our control system
would become so sluggish or so poorly coordinated due to slow response speeds
and general massiveness of the balloon system itself that the gas burner and
exhaust systems could become out of phase with their responses to updrafts or
downdrafts? Obviously, it would do us no good to have a system that, because of
cheap components or poor judgment in hardware design, responded so slowly to
command signals that when the balloon calls for an increase in altitude, the vent¬
ing valve on the air bag is still open.
As we soon find out in our study of automatic control systems, the behavior
or response of the automatic altitude system of our hot air balloon is very typical
of other automatic control systems in general. And to comment on the question
raised in the previous paragraph, out-of-phase response is a serious problem in
automatic control design.

SUMMARY

In studying the audio amplifier, such as the ones typically used in hi-fi systems, we
developed an understanding of frequency response, gain, bandwidth, and feed¬
back. Even though all of these terms were applied to the audio amplifier, they are
common terms applied to automatic control systems in general.
* Another important concept in automatic control design is phase. In the case
of the amplifier, we found that by reducing or eliminating an out-of-phase signal
response problem, we could increase the bandwidth of our amplifier. We found
that this phasing problem could be affected through the introduction of negative
feedback.
72 Chapter 2 Control Theory Basics

Positive feedback has little if any use in the design of an automatic control
system; however, it is vital in the design of electronic oscillators. On the other
hand, negative feedback is vital to the design of automatic control systems.
The following important equations were discussed in Chapter 2:

Eo
Gain (Eq. 2-2)
Em

Eo A
For positive feedback: (Eq. 2-8)
E\n 1 - AB

E0 A
For negative feedback: (Eq. 2-14)
Em 1 + AB

the system’s output


Transfer Function (Eq. 2-15)
the system’s input

EXERCISES

2-1. If an automatic control system has an input voltage of 15 volts, an amplifier


gain of 3, and a B value of 0.25, find its output voltage.
2-2. Determine the amount of feedback needed to create a smoothly running
oscillator having an amplifier whose gain is 25.
2-3. Design an automatic control system using a block diagram style (similar to
Figure 2-15) showing how a temperature controller would be used to regu¬
late the temperature inside a refrigerated chamber used for research applica¬
tions. There are several possible methods that can be used for this design.
2-4. Write the transfer function for a gear box transmission whose output shaft
rotates 100 revolutions for every 726 revolutions of its input shaft.
2-5. What is the transfer function for a hydraulic cylinder that produces a piston
rod displacement of 49 inches with a supply of hydraulic fluid of 2.37 gal¬
lons?
2-6. Determine the gain of an amplifier needed in an automatic control system to
produce an overall system gain of 0.9, assuming a feedback value of 0.37.
2-7. Calculate the decibel output of an amplifier having a gain of (a) 1,000
(b) -21.5 (c) 14.5 (d) 0.65.
2-8. Determine the output voltage resulting from a system having a gain of 135
dB and an input voltage of 7 VDC.
2-9. Show mathematically that a 3-dB down point is the same as a voltage gain of
0.707.
2-10. For a negative feedback system, how much feedback is needed in order to
create an overall gain of 0.25 for a system assuming an amplifier gain of 20?
References 73

REFERENCES

DeRoy, Benjamin E., Automatic Control Theory, New York, N.Y.: John Wiley
& Sons, Inc., 1966.
Miller, Richard W., Servomechanism Devices and Fundamentals, Reston,
Va.: Reston Publishing Co., 1977.
Sante, Daniel P., Automatic Control System Technology, Englewood Cliffs,
N.J.: Prentice-Hall, 1980.
3-1 THE MATHEMATICS OF AN AUTOMATIC CONTROL
SYSTEM

Perhaps a better title for this chapter would have been, “Laplace Transforms—
The Painless Way of Avoiding Calculus Problems.” In working with automatic
control systems, it is often necessary to describe their operations using rather
complicated differential equations and calculus expressions. Needless to say, for
many engineers and technicians, this can be a rather somewhat laborious ordeal.
Unfortunately, though, some of these math approaches defy any other methods of
solving, and so in some cases no choice is left but to use calculus. For someone
just starting to understand a new engineering concept, the beauty and enjoyment
of understanding and working with that concept or system can be easily destroyed
by becoming bogged down in the mathematical justifications. At one time or
another we have all become victims of this problem.

3-2 A LITTLE BACKGROUND CONCERNING LAPLACE


TRANSFORMS

About 200 years ago, Pierre Simon, Marquis de Laplace (pronounced Mar-kee
Day Laplahz), developed a method of greatly simplifying the solutions of differen¬
tial equations. He must have had the modern-day controls engineer in mind.
(Actually, it was Oliver Heaviside who, some years later, really perfected the
modern-day principles of Laplace transforms. But that’s another story.)

74
Section 3-3 How Transforms Are Used 75

3-3 HOW TRANSFORMS ARE USED

Let’s see how Laplace transforms can make life so much easier for us. Think of
performing the division problem, 477,123 -r- 147, in your head. A difficult task,
right? But what if, instead, we were to present the same problem in this form: 1 x
105.67863 j x |q2 16732 Almost immediately, we see the answer as being 1 x
103-51131. Simply by using the laws of scientific notation, we arrive at our answer.
However, to get our answer into a more familiar form, we have to determine the
antilog of 3.51131, which turns out to be 3245.7. Admittedly, this job requires
using log tables or a calculator.
In using Laplace transforms, the general idea is somewhat the same. You
must first convert a given differential equation expression to its Laplace equiva¬
lent. Then, you would do whatever algebraic manipulations would be necessary
to arrive at your solution, then reconvert the expression by converting the anti-
Laplace of that expression back to its original form. Basically, you will have
reduced what could have been a rather complicated calculus solution down to
some very simple algebraic steps.
Laplace transforms are used in time-dependent types of expressions, that is,
expressions that say y = f{t). More specifically, they can be used only for solving
linear differential equations whose variables are functions of time. In automatic
control expressions these will, fortunately for us, occur more frequently than not.
We won’t need differential equations for the work we do here. Instead, we
look for certain time-related math expressions that crop up periodically in the
course of performing math operations. When we spot these expressions, we go to
a lookup table of Laplace transform equivalents and make the appropriate substi¬
tution. After doing that, and after performing what algebra is necessary to arrive
at the form that we want, we then look up the inverse transform to get the
expression back into the time domain.
Let’s discuss an example here to see how this works. Look back at Figure
1-32 in Chapter 1. In this figure we see a capacitor, C, (assumed initially to be
uncharged) in series with a resistor, R. Between terminals A-B, we apply a step
voltage e(t) (i.e., a voltage whose magnitude rises to a particular value immedi¬
ately at t = 0). By using differential and integral calculus, we can arrive at the fact
that the current, i, = (E/R)(e~tlRC). Now note the expression e~t,RC in the expres¬
sion just mentioned. This occurs time and again throughout discussions of sys¬
tems that have transient conditions. Remember, transient conditions last only for
a small period of time and then disappear. Many math operations in automatic
controls involve this particular expression. As a result, things can get fairly
messy during the course of performing these operations, especially when using
calculus. Fortunately, this expression has been assigned a Laplace transform
because of its frequent use.
The best way to appreciate all of this and to get a better foothold in under¬
standing all that has been said so far is to work an example problem. Again, look
at Figure 1-32. Let’s say that we want to develop a mathematical expression that
76 Chapter 3 Laplace Transforms

relates the capacitor’s (C) charging current, i(t), to the capacitor’s time constant
(RC). In other words, we want an expression that says: i(t) = f(RC). To do this,
we have to find what f(RC) is.
First, we look at the involved solution. If you have no interest whatever in
this solution, which will require integral and differential calculus, you can skip this
part and go on to the simpler solution. Otherwise, here now is our involved
solution:

EXAMPLE -
Involved Solution:

The voltage, e(t) = RC + vc(t) (3-1)


dt

Let e{t)

dvc(t)
then E = RC ■d + vc(t) (3-2)

We want to find vc(t) because vc = E - iR from Kirchhoff’s voltage law. And from
this expression we can solve for i, which is:

vr
i —
R

which can be restated as


vc{t)
i{t) = (3-3)
R

Now, to find vc(t) so that we can place it into Eq. (3-3) for a solution of i(t), we want
to set up Eq. (3-2) with vc(t) on one side.
dvc{t)
E - vc(t) = RC
dt

dvc(t) _ 1
And, dt
E - uc(t) RC

Integrating both sides:


dvc(t) 1
dt
E - vc(t) RC

-t
or Ln(E - vc(t)) + K
RC

where K is a constant of integration.


When t — 0, then LnE = K. Substituting K = LnE into Eq. (3-3) will give:

Ln{E - vc(t)) = + LnE


Section 3-3 How Transforms Are Used 11

-t
or Ln(E — vc(t)) — LnE =
RC

, E - vc(t) -t
or Ln
RC

Taking the natural antilogs of both sides results in:

E - vc{t) — e-tlRC

Solving for uc(r),


Kn-tIRC
E — vc(t) = Ee —

Solving for vc(t),

vc(t) = E - Ee~tlRC
= E( 1 - e~tlRC)

Since i(t) = E ~fl) (Eq. 3-3),


R

E - E{ 1 - e~tlRC)
then i(t) =
R

E(e~tlRC)
and finally, /(t) = (3-4)
R

Now, the simpler solution for the same problem:

EXAMPLE
Simpler Solution:
Remember, we are trying to find the expression for i(t), the charging current for
capacitor, C. First, we note that the reactance for C is:

x = _!_ or _L
c 2irfC jo>C

We also know that ZT (the total circuit impedance as measured across terminals A-B
in Figure 1-32) is

1
Zj — R +
joC

Going to a table of Laplace transforms (Table 3-1), we see that ja> has a transform, 5.
Therefore,

ZT — R + — (3-5)
78 Chapter 3 Laplace Transforms

According to this same table, E (our voltage step) = E/s. And since E = IZ, we can
therefore state that E(s), which is E/s in our case, = I(s) x Z(s). This means that
E/s = a Laplace transform expressing impedance as a function of In other words:

E(s)
Hs) =
Z(s)
and since E(s) = E/s and, referring to Eq. (3-5), Z(s) - R + 1 /sC, then

E E
s s = EC
I(s) =
sRc + 1 (sRC + 1)
R + f
sC lc
Now, multiply the far right equation by R/R, which of course is 1, to get:

R EC ERC
R ' sRC + 1 - R(sRC + 1)

Since RC = r, then

Er
/(s) (3-6)
R(st + 1)

Now, to find the time equivalent of Eq. (3-6), we have to determine its in¬
verse transform. Referring again to Table 3-1, we find from pair 6 that the inverse of
1/(jt + 1) is (\/r)e tlT. Notice that Er/R is a constant. Constants multiplied by time
functions are not changed by transformations, and vice versa. And since I(s) trans¬
forms back to i(t), we finally have:

Er 1 — th
m
R r e

or (3-7)

Equation (3-7) checks with Eq. (3-4), which was obtained using the involved method.

Admittedly, at this point it’s probably difficult to see how using Laplace
transforms is any less complicated or involved than using the calculus approach.
But, once you become familiar with the Laplace method, you’ll begin to see its
benefits.

3-4 TRANSFORM TABLES

Now that we have demonstrated the using of transforms showing the possibility of
making our math somewhat simpler, let’s take a closer look at the other trans¬
forms listed in Table 3-1. You will probably find a number of familiar looking
expressions that will have a not-so-familiar Laplace equivalent transform. After a
Section 3-4 Transform Tables 79

TABLE 3-1 LAPLACE TRANSFORM PAIRS

Pair
no. m F(s) Comments

/ E
1. I or E A step input of magnitude I or E.
s’ 5
1
2. 1 A unit step input.

A
3. At A ramp function.
s2
2A
4. At2 A parabola.
s3
A
5. Ae~at
s + a
Ae~th A General response of first-order
6.
T TS + 1 system.
A First-order response to a step
7. A(1 - e~'h)
s(ts + 1) input.
A(1 - e~at) A
8.
a s(s + a)
Aco
9. Asincot A sine function.
S2 + CO2

As
10. Acoscot A cosine function.
S2 + CO2

11. CO sd

12. jco s
13. a s20

14. - Z2)
vV? sm(a,"Vl

Aco2 Second-order system, general


s2 + 2 zcons + co2 response where z < 1.
Ate~t/T A General response of second-
15.
T2 (TS + l)2 order system (z = 1).
A
16. Ate~at
(s + a)2
g~ZOJnt /
17. Vl - z2\
Vl — z2 t — Arc tan
1 + yj _ z2 sin v"'1 —z '
co2 Second-order system response to
' s(s 2 + 2 zcons + co2 a unit step input (z < 1).

a/ 1 — z2
where 0° < Arc tan-< 180°
—z
80 Chapter 3 Laplace Transforms

while, however, you will feel just as at home with the Laplace equivalent as you
will with the now more familiar time domain expressions.
Because automatic controls deal with system characteristics that can be
short-lived or transient in nature, the mathematical expressions for these transient
conditions are seen as decaying functions. The charging capacitor is a good exam¬
ple. The equation for this charging condition is seen in Figure 1-32. As you look
through the transforms in Table 3-1 and their time domain counterparts, notice
how often the expression, e~Kt, occurs. This should tell you how important that
particular exponential function is to automatic control systems and how often it
occurs.
The derivations of all the Laplace transforms follow three basic steps:

1. Choose the time domain expression that you want to convert to a Laplace
transform. In other words, this expression must be in the form of y = f(t).
2. Multiply this expression by a factor, e~st. This is a factor that decreases or
decays to zero as time becomes infinitely large.
3. Integrate this new expression with respect to time between the limits of zero
and infinity.

Mathematically, these three steps can be expressed in the following way:

F(s) = £[/(/)] - f“ f(t)e~s‘ • dt

where F(s) = the Laplace transform expression


£[/(£)] = to be read as “take the Laplace transform of the time domain
expression. . .
= — to be read as “defined as. . . .” or “the same thing as. . . .”

But again, we really won’t have to concern ourselves with the preceding
three rules, or with the preceding mathematical expression for deriving our trans¬
forms. This has already been done for us. The information has been given here
only to complete our discussion of transforms. Hopefully, you may become inter¬
ested in the mathematical proofs at a later time, since there is no better way to
understand this material. If so, the seed has been planted here for that study.
While we are defining the various Laplace math symbols that we use soon,
another symbol is £-1[F(s)]. This means, “take the inverse transform of the
Laplace expression. ...” Therefore,
£-'[F(s)] s/(/)

Table 3-2 shows the various Laplace mathematical operations and their time
domain equivalents. Looking at this table, the only differences in the math opera¬
tions in the two systems are differentiation and integration. Since we plan not to
do either, we won’t worry about them. These two operations are presented here
only to offer a complete table of math operation equivalents.
Section 3-5 Examples Using Transforms 81

TABLE 3-2 EQUIVALENT MATH OPERATIONS USING


LAPLACE TRANSFORMS

Math operation m Fis)

Addition of two functions fi(t) +• flit) F\is) + F2(s)


Subtraction of two functions flit) ~ flit) Fiis) - F2(s)
Multiplying a function by a constant flit) • K Fis) • K
Multiplying two functions together flit) ■ flit) F^is) ■ F2(s)

flit) fis)
Dividing a function by a constant
K K
m fis)
Dividing two functions
fiit) fiis)
df(t)
Differentiation s • F(s)
dt

Integration jf«)i dt — ' F(s) •


5

3-5 EXAMPLES USING TRANSFORMS

The only good way to fully understand and appreciate the power of the Laplace
transform is to become involved in some controls-type problems. At this point,
however, we still need to digest some additional theory. So, instead, let’s look at
some basic problems involving transforming an expression from one domain to
the other.

EXAMPLE 3-1 -
Transform/(r) = 25 to a Laplace transform. Note that 25 could be a step input to a
circuit having a magnitude of 25 volts or 25 amps, etc.

Solution:
Referring to Table 3-1, we see that pair 1 appears to fit the description of the given
problem. Therefore,

EXAMPLE 3-2
Transform 1/(5 + 5) into an f(t) expression.

Solution:
1
Fis) =
is + 5)
82 Chapter 3 Laplace Transforms

Referring to Table 3-1, we see that pair 5 appears to fit the wanted description, where
A = 1 and a = 5. Therefore,

m = (l)e-5‘

EXAMPLE 3-3
Find £[/(/)] = 10 +

Solution:
Noting that this is an addition problem of two F(s) expressions, we have to refer to
Table 3-2 to see that/,(7) + f2(t) = Ffs) + F2(s). Therefore, referring to pairs 1 and 5,
we have:

rrcvvi 10 ^ 1
£[/W] " 5 + 5 + 1

EXAMPLE 3-4
Find £~l[4/{s2 + 16)].

Solution:
Look at pair 9 in Table 3-1. (We discarded pair 10 as a possibility since there was no
5-term in the given function’s numerator.) Looking at the given function, we note
that the co-term in the denominator is a perfect square of the co in the numerator. This
is what pair 9 requires. Also, we can say that A - 1. Therefore,

£_1[F(5)] = sinAt

EXAMPLE 3-5 -
Find F{s) for the function f{t) — 121.

Solution:
Pair 3 shows that for/(r) = A • t, F(s) = A/s2. Therefore, with A = 12, F(s) = 12Is2

EXAMPLE 3-6 -
Find F(s) for the function f(t) = 50(1 - e-100')-

Solution:
Look at pair 8. Notice that A appears to be 50 in the given function, but we must
have an a under the 50 (which is 100 in our function) in order to use this pair. So, our
Summary 83

only recourse is to force a 100 under the 50. But to do this and not change the value
of the function itself, we must multiply the function by 100. Therefore, the result is:

100
fit) = 50(1 - e~mt) = x 50(1 ,-ioon
100
5,000 100r
(1 -
100

We now see that A actually = 5,000 and a = 100, and we can now proceed with using
the F(s) conversion given in pair 8, which is:

5,000
s(s + 100)

EXAMPLE 3-7 --
Given that f(t) = (5t/36)e~tl6, find F(s).

Solution:
Look at pair 15. Notice that the 6 in the given exponent is a square root of the figure
in the denominator of At!36. Therefore, let 6 = r. Also, let A = 5. Then, F(s) =
5/(65 + l)2.

SUMMARY

Laplace transforms represent a straightforward alternative mathematical ap¬


proach to automatic control system analysis. (We investigate this further in the
coming chapters.) The Laplace method allows us to convert a time domain /(/)
expression into the so-called 5-plane domain or f is) expression, and then back
again. Table 3-1 lists the most commonly used conversions that are used in this
book.

EXERCISES

Given the function fit), find F{s) using Tables 3-1 and 3-2.
3-1. fit) = 107
3-2. fit) = 2012
3-3. fit) = It1 + 9t - 2

3-4. f{t) = § 0 - e-10')

3-5. f(t) = 2.5(1 - e~m)


84 Chapter 3 Laplace Transforms

3-6. /(f) = 2t2 + 6(1 - e-34')


3-7. /(/) = 5e~tl4
3-8. /(/) = l.4e~Ut
3-9. /(/) = (1 - e-<«)
3-10. f(t) = (2.9e~5t)(2e~tl4)
3-11. fit) = 3.2sina>t
3-12. f{t) - 5cosl44r Hint: What does o = ?
3-13. /« = 2te-°-5f
3-14. t{t) = %ie~tl6 - e~tl]2)
In the following expressions, find fit) for the given Fis) functions:
22
3-15. Fis) =
is + l)2

3-16. Fis) = /r. ^—;—7- Hint: What is the value of r?


5(0.025 + 1)

6.2
3-17. Fis) =
5 + 3
23.5oj
3-18. Fis) =
52 + co3

3-19. F(s) s2 + 1>00()

3-20. Using pair 14, determine the value of a>n and z in the following expression:

3.96
Fis) -
0.252 + 5 + 3.96

REFERENCES

Kuhfittig, Peter K. F., Basic Technical Mathematics with Calculus, pp. 921-34.
Monterey, Calif.: Brooks/Cole Publishing Co., 1984.
Rice, Bernard J. and Jerry D. Strange, Technical Mathematics and Calculus,
pp. 827-38. Boston, Mass. Prindle, Weber & Schmidt Pub., 1983.
Washington, Allyn J., Basic Technical Mathematics with Calculus, (4th. ed),
pp. 933-45. Menlo Park, Calif.: The Benjamin/Cummings Publishing Co.,
1985.
Transducers and
Control System
Components

4-1 DEFINING TRANSDUCERS

In general, a transducer is a device that transforms one form of energy into


another. A transducer is generally thought of as a device that is relatively small;
that is, it can usually be held in the hand, and the amount of energy being trans¬
formed is usually quite small. In addition, the conversion process between the
input and output is done quantitatively. That is, by knowing the input quantity,
one can predict the amount of output based on a predetermined calibration factor,
or set of factors. (A furnace is not usually considered a transducer even though it
is an energy converter. This is because of its sheer size and because its output has
not been precisely correlated with its input as in the case of a transducer.) The
kinds of transducer devices that we discuss here can be hand-held for the most
part. They can change quantities such as heat, light, motion, etc. into a form of
electrical energy. More specifically, each transducer is specifically designed to
change a particular quantity into a small current or voltage signal that will be
proportional in magnitude in some way to the quantity being sensed.
In automatic control systems, transducers are used extensively to sense
position, motion, or some other quantity. This information is then transmitted in
the form of an electrically equivalent signal to some sort of control circuitry for
the purpose of producing the desired response (Figure 4-1).
All transducers contain some sort of sensing element or device. This is the
portion of a transducer that actually does the transducing. Figure 4-2 shows,
schematically, the difference between the sensing element of a transducer and the
transducer itself. It is important to understand the difference between these two

85
86 Chapter 4 Transducers and Control System Components

Figure 4-1 A position controller.

terms. Each has its own particular transfer function. The units of measurement
may be the same for both, but the magnitude of the transfer function may be
radically different. And it is the transducer’s transfer function that we are primar¬
ily interested in. Admittedly, up to this point we have developed little under¬
standing of the importance of transfer functions in automatic controls, much less
developed an understanding of what an automatic control circuit is. This will
come later. For the present, however, we must become familiar with the transfer
functions of the various devices discussed in this chapter, and then, as we get into
the heart of automatic control theory, the purpose of this type of function will
become clear.

Transducer

Figure 4-2 The circuit of a typical transducer.


Section 4-2 Position Sensing 87

4-2 POSITION SENSING

An example of position sensing in an automatic control system is found in the


application of positioning a radar antenna. The operator, sitting inside a control
bunker, has the ability to remotely control the positioning of his or her radar
dish. If a gust of wind comes along to disturb that position, the dish’s rotation
control circuitry (containing positional transducers in the dish’s mount) will sense
the disturbance and counteract the movement by causing the dish to move back to
its original position. This is done automatically without the operator’s interven¬
tion (Figure 4-3).

All of This Located in a Remote Area away from Radar Dish


I-1

Figure 4-3 A radar’s antenna positional control system.

4-2.1 The Potentiometer

One of the most frequently used devices used for sensing position is the potenti¬
ometer. Potentiometers, or “pots,” are also called variable resistors. When used
for automatic control system applications, these pots are generally of the preci¬
sion type. That is, the resistive element inside the pot is constructed of a durable
temperature-stable material that has been formulated to create the same resistive
values over a long period of time. The resistances measured at the pot’s output
terminals are usually very repeatable to within a very small percentage.
' There are two major types of pot configurations used in automatic control
circuits. These are shown in schematic form in Figure 4-4, (A) and (B). Schemat¬
ics (C) and (D) in that same figure are the actual schematic representations used
for either configuration shown in (A) and (B). However, (D) is preferred. Notice
that (C) doesn’t show any output connection for the wiper. This information is
88 Chapter 4 Transducers and Control System Components

Resistive
Element

Resistive
Element

Figure 4-4 Potentiometer configurations.


A. Linear pot
B. Rotary pot
C. Electrical symbol
D. Alternate electrical symbol

vital in some automatic control system circuitry. Figures 4-5 and 4-6 illustrate
typical linear-constructed and circular-constructed pots.
Referring again to Figure 4-4, let’s now analyze the pot in order to determine
its transfer function. Looking at (A), we can see that the input quantity to the
wiper is a displacement. This is depicted in the wiper’s drawing by the label wiper

Figure 4-5 A linear precision pot. (Courtesy of Waters Manufacturing, Inc.


Way land, Mass.)
Section 4-2 Position Sensing 89
■#«£$

Figure 4-6 A circular precision pot


and transducer. (Courtesy of Waters
Manufacturing, Inc., Wayland, Mass.)

motion. The output from the pot is a variable voltage, labeled E0ut, and is mea¬
sured between the wiper and one of the two ends of the resistive element.
Consequently, this pot’s transfer function would be:

£out = K volts
(4-1)
/ p inch

where Kp = gain or transfer function value of pot (volts/in.)


Eout = output voltage (AC or DC volts)
/ = displacement of wiper (inches, or any other convenient unit of dis¬
placement)

EXAMPLE 4-1 -
Find the value of Kp for a linear pot whose output shaft must be moved a distance of
5.87 inches in order to create an output voltage of 12 volts DC for a positioning
circuit.

Solution:
According to Eq. (4-1), the pot’s transfer function value is calculated by the following
method:

12 VDC
Kp 5.87 in.
- 2.044 volts/in.

The transfer function for the circular pot (Figure 4-4, Schematic B) is calcu¬
lated similarly. In this case, however, since the input displacement to the wiper is
circular, the Kp value would be:

E0Ut Tr VOltS
(4-2)
d deg
90 Chapter 4 Transducers and Control System Components

where Kp = gain or transfer function value of pot (volts/0)


E0ut = output voltage (AC or DC volts)
6 = angular displacement of wiper (deg. or rad.)

There are other types of transducers other than the pot that can be used for
sensing displacement or position. The pot, however, is probably the most fre¬
quently used and is the least expensive of all devices. As you become familiar
with other types of transducers, you will be able to see other methods of detecting
position other than just the one mentioned here. Refer also to the discussions on
electromagnetic sensing (Section 4-7), light sensing (Section 4-9), and resolvers
(Section 4-13).
One of the greatest disadvantages in using pots for position sensing is that
their mechanical construction frequently prohibits them from rotating a full 360°.
This can be a serious restriction in some applications using an automatic control
design. In addition, pots tend to wear out more quickly than do some other types
of position transducers.

4-2.2 The LVDT

LVDT stands for linear variable differential transformer. This type of transducer
is used for sensing position and is noted for its precision and accuracy for perform¬
ing this function.
The LVDT works on the principle of transformer action. The location of a
moveable plunger, acting as a transformer core, determines the amount or magni¬
tude of signal coupled to the two secondary windings of the LVDT. Also, the
plunger’s position determines the phase of the secondary signal. This phase is
determined relative to the primary signal. The secondary signal may either be in
phase for one direction of travel, or it may be shifted 180° out of phase for an
opposite travel direction, depending on the core’s relative position with its center
of travel. The transfer function for an LVDT is:

Lout _ „ volts
(4-3)
/in KlVDT inch

where KLVDT the sensor’s gain or transfer function value for LVDT (volts/in.)
E,out output voltage (either AC or DC volts)
I'm input displacement of core (inches)

4-2.3 The Linear and Rotary Photocell Encoder

The linear photocell encoder consists of a photocell and light source that has a
transparent strip or disk with opaque closely spaced bars which are passed be¬
tween the photocell and light source. Displacements as small as 0.0001 inch can
be detected. The transfer functions for these devices are similar to those of the
linear and rotary pots described in Section 4-2. (See equations 4-1 and 4-2.)
Section 4-3 Velocity Sensing 91

4-3 VELOCITY SENSING

Many control systems require having to know the linear or rotational velocity of
certain components for the system’s proper operation. An example might be the
cruise control system on an automobile as it travels down a highway. Once the
desired speed of the car is set by the driver, the system must be able to measure
the car’s forward velocity in order to make corrections due to wind velocity
changes and changes in the slope of the highway.
There are numerous methods available for measuring velocity. We discuss a
few methods here and, after having done this, you will undoubtedly see other
ways of making this kind of determination.
Quite often, it is possible to use a positional type of transducer, such as the
precision pot discussed earlier, and by making a positional measurement during a
measured length of time, the velocity becomes known. Let’s investigate an exam¬
ple here. Refer to Figure 4-7. We see a linear measuring pot being used, similar to
the ones shown in Figure 4-5, to measure the velocity of a rod and piston on an
hydraulic cylinder. The change of DC voltage per unit of time can be directly
correlated to the input velocity of the cylinder’s follower attached to the pot’s
wiper arm. This is the purpose of the time base in Figure 4-7. This is an electronic
circuit that produces a signal that is proportional to the distance traveled per unit
time. A clocking circuit periodically samples the distance voltage coming from
the linear pot, each sample being measured over a precisely known length of
time. The output of the time base circuitry may be in the form of a variable
amplitude DC voltage, or it could be a variable frequency AC waveform. The
transfer function for this transducer device would then be:

Lout= K volts
(4-4)
vel p ft/sec

where Kp = sensor constant or transfer function value for resistive-type velocity


sensor. Must be obtained from manufacturer or experimentally de¬
rived.
Eout = output voltage of transducer (AC or DC volts)
vel — input velocity of wiper arm (ft/sec)

Figure 4-8 shows a typical velocity measuring transducer used for determin¬
ing linear velocity. It can also be used for rotational velocity when used with the
proper mechanical rotary-to-linear motion conversion techniques, as is often
done. The sensor that it uses is a photocell device that detects the passage of
light- and dark-spaced marks on a rotating disk. The photocell in turn produces
either a voltage or no voltage signal, depending on the light level that it senses,
and sends this pulsating voltage level to a signal conditioning circuit. This circuit
then converts the frequency of the pulses into a proportionally varying DC voltage
level whose magnitude represents the velocity amount. The transfer function for
this transducer would be determined by:
92 Chapter 4 Transducers and Control System Components

Figure 4-7 Measuring velocity of a hydraulic cylinder using a linear pot. (Photo
Courtesy of Waters Manufacturing, Inc., Wayland, Mass.)

j?
^out Kphoto volts
(for a linear transducer)
vel ts + 1 ft/sec
volts volts
or (for a rotary encoder) (4-5)
rad/sec deg/sec

where Kpholo — sensor gain or transfer function value for velocity photocell trans¬
ducer. (volts/ft/sec, volts/rad/sec, or volts/deg/sec)
■£"out = voltage output (DCV)
vel = input velocity (ft/sec, rad/sec, or volts/deg/sec)
Section 4-3 Velocity Sensing 93

— = Velocity
At

Wh
\ i

DC Voltage Frequency-to- Voltage


Output Converter
Photocell
Detector

Figure 4-8 A velocity sensing transducer.

5 = Laplace transform for jo)


t— time constant of photocell (sec)

Note that in the preceding transfer function you see the Laplace transformation
expression, (rs + 1). This is a transform associated with many sensing elements
or transducers whose outputs are not instantaneous, but rather gradual in that
they require time to build up their outputs to their final values. The characteristic
of this response is similar to the R-C time constant discussed in Chapter 1. The
mathematical expression of this characteristic is the familiar e~t,T function seen
repeatedly in the Laplace transform Table 3-1 in Chapter 3. The theory of opera¬
tion of the photocell is explained later in Section 4-9.

4-3.1 The Rotary Magnetic Encoder

The circuit in Figure 4-8 can be modified to use a magnetic sensing device to
detect rotation of the disk rather than using a photocell. The disk itself would of
course have to be modified so that the magnetic sensor could sense its degree of
rotation. Often, what is done is to use a many-toothed gear made of ferromagnetic
material such as iron. The magnetic sensor, sensing a change in the magnetic field
that it creates, as the result of a tooth or any other protrusion sweeping past it,
generates a pulse of DC voltage. The frequency of this pulse would of course
depend on the rotational velocity of the disk or gear. As in the case of the photo¬
detector circuit in Figure 4-8, the frequency of the pulses could then be converted
to a variable DC voltage output. The transfer function for this system would be
identical to that of the photo-detector device, except perhaps for the magnitude of
the constant, K.
Another common method used for measuring velocity is the rate generator.
This is basically nothing more than a DC voltage generator whose output voltage
is directly proportional to the input velocity of its rotating shaft. The principle of
94 Chapter 4 Transducers and Control System Components

operation and the method of calculating its transfer function is discussed in detail
in Section 4-11.

4-4 PRESSURE SENSING

The sensing of fluid pressure, whether that fluid is a liquid or a gas, can be done
using similar transducer methods. Perhaps the most commonly used method for
sensing pressure is through the usage of strain gages. The principle of operation
of a strain gage rests in the fact that if one were to stretch an electrical conductor
such as a wire, that wire’s DC resistance would increase proportionally to the
applied stress. Very small changes in the wire’s resistance can be detected by
arranging the wire in a Wheatstone Bridge configuration, seen in Figure 4-9. The
resistor, R3, represents the stressed wire, whereas the other three resistors, R],
R2, and R4 are precisely known and temperature-stable resistors. Often, R4 is
another strain gage identical to the strain gage at R\. However, unlike R\,R4 does
not receive any stress. Its sole purpose is to compensate, that is, nullify, any
changes in temperature occurring to the stressed strain gage. It can do this be¬
cause of the current-balancing characteristics of the Wheatstone Bridge and the
locations within the bridge of R\ and R4. Being sensitive to temperature is one of
the unfortunate characteristics of a strain gage. The unstressed strain gage at R4 is
called a dummy gage.
A voltage source, V, supplies a voltage to the bridge. By design, when R3 is
unstressed by no pressure being present on the transducer’s pressure diaphragm,

Applied
Pressure

Figure 4-9 A pressure sensing trans¬


ducer.
Section 4-5 Sound Sensing 95

the opposing currents flowing through the bridge’s central leg are equal and are
therefore nullified. Consequently, no current at all flows under this condition.
However, as soon as a pressure occurs at the transducer’s diaphragm, R3 becomes
stressed, thereby upsetting the balanced condition of the bridge and causing cur¬
rent to flow through the indicator in the bridge’s central leg. The amount of
current will be proportional to the amount of pressure being applied. This result¬
ing imbalance of current is then transmitted to a signal conditioning circuit that
converts the current into a designed signal form such as a variable DC voltage.
The transducer’s transfer function becomes:

£out _ Kpr VOltS .


Pm 1 + st lbs/in2

where Kpr sensor constant or transfer function value for pressure transducer.
Must be obtained from the manufacturer or experimentally ob¬
tained.
output voltage of transducer (DCV)
input pressure to transducer (lbs/in2)
Laplace transform for joj
system time constant (sec)

4-5 SOUND SENSING

The sensing of sound involves using one of the more frequently used transducers,
as in the use of the microphone. Admittedly, this form of transducing is not all
that often used in automatic control applications, but nevertheless, it is discussed
here to round out the discussion on transducers.
Before we begin this discussion, however, we first must understand what
sound waves are and how their intensities are measured. Sound waves are noth¬
ing more than pressure fronts emitted from a sound source. The pitch of the
sound is determined by the frequency of the sound waves, that is, the number of
pressure fronts passing a particular point each second of time. The sound’s inten¬
sity is measured in units of the decibel.
There are many ways to convert sounds into electrical signals, so we start
with one of the simplest and earliest methods used in sound sensing. Figure 4-10
shows the construction of the carbon-type microphone. This was the earliest type
of device used to convert sound into electrical signals. The principle of operation
is quite simple and ingenious. A diaphragm was fastened rigidly to a flexible
container holding dried carbon granules. The diaphragm was free to vibrate
whenever struck with sound waves. This vibration was passed on to the carbon
granule container where the granules would be alternately compressed and
96 Chapter 4 Transducers and Control System Components

Figure 4-10 The carbon button micro¬


Load phone.

allowed to decompress with each impinging wave front on the diaphragm. A small
electrical signal was passed through the granules so that the electrical current
would experience a proportional increase and decrease in electrical resistance
depending on whether the granules were being compressed or decompressed at
that instant. This would result in a varying voltage or current that would result in
a fair representation of the frequency and intensity of the sound picked up by the
microphone.
Figure 4-11 shows another type of microphone. This one uses a piezoelec¬
tric substance (see Table 4-1) to convert sound waves into a varying output volt-

Figure 4-11 The piezoelectric or


“crystal” microphone.

age. When a piezoelectric substance experiences an applied force, an output


voltage is generated. The sensitivity (i.e., the ability of the substance to generate
a voltage for a given input of force) is often stated in coulombs/newton, the
coulomb being a unit of electrical charge, whereas the newton is a unit of force. In
the case of a piezoelectric microphone, sound waves are allowed to strike a
diaphragm that alternately compresses and decompresses an attached piezoelec¬
tric crystal. Table 4-1 shows the varying output sensitivities of various and often-
used materials in microphone construction.
Figures 4-12 and 4-13 show two additional microphone types that both per¬
form very much the same way as the one just described, but utilizing different
Section 4-5 Sound Sensing 97

TABLE 4-1 COMMONLY USED PIEZOELECTRIC


MATERIALS1

Sensitivity
Material type (coulomb/newton)

Quartz 2.3 x 10“12


Tourmaline 1.9 x 1012
Rochelle salts 550.0 x 10'12
Ammonium dihydrogen phosphate 48.0 x 10-12
Lithium sulfate 16.0 x 10“12

1 Richard L. Allen and Bob R. Hunter, Transducers, (Al¬


bany, N.Y.: Delmar Publishers, 1972), p. 35.

processes. The microphone in Figure 4-12 contains a capacitor whose one plate is
attached to the vibrating diaphragm. This creates a varying capacitance between
the two plates. A voltage is supplied to the capacitor to allow the charge to vary.

Figure 4-12 The capacitive micro¬


phone.

On the other hand, Figure 4-13 shows a microphone that generates a proportional
voltage because of a magnet and coil combination. A magnet is attached to the
diaphragm and is allowed to vibrate back and forth inside a coil. The lines of
magnetic flux sweeping past the coiled turns then generate a voltage that is pro¬
portional to the amount of vibration. This voltage, in turn, will be proportional to
the arriving sound waves at the microphone.
One of the more recent innovations utilizing sound waves in an automatic
control circuit is the automatic focusing unit used in certain instant cameras.

Figure 4-13 The “dynamic” or elec¬


tromagnetic microphone.
98 Chapter 4 Transducers and Control System Components

Figure 4-14 shows this device along with its circuitry. An ultra-high-frequency
sound pulse is radiated from the camera’s sound transmitting circuitry toward the
subject. An echo is detected by a sensitive microphone located back on the
camera and, after timing the interval between pulses, the camera’s circuit com¬
putes the subject’s distance and then automatically adjusts the camera’s optics for
that distance.
The one thing that all these microphones have in common with each other,
aside from the fact that they all have the capability to sense sound, is that their
transfer functions are all alike. Their transfer function is:

K sonic
(4-7)
1 + ST

where Adonic = sound detector constant or transfer function value. Must be ob¬
tained from manufacturer or experimentally obtained.

Figure 4-14 Camera with automatic sonic focusing. (Courtesy of Polaroid Cor¬
poration, Cambridge, Mass.)
Section 4-6 Flowrate Sensing 99

£"0ut — DC output voltage (volts)


PdB = input sound pressure level (dB)
5 = Laplace transform for jcn
t = Sound detector’s time constant (sec)

4-6 FLOWRATE SENSING

There are many schemes that are used for sensing the flow-rates of gases and
liquids. A common method is to immerse a paddle-wheel-like mechanism into the
fluid and to allow the flowing liquid to turn the paddle wheel much like a water
wheel on a mill house or a windmill. These devices are surprisingly accurate.
Figure 4-15 shows one such device along with its schematic. Notice that it uses a

Figure 4-15 A flowrate transducer for sensing liquid flows in a pipe. (Courtesy of
Flow Technology, Inc., Phoenix, Ariz.)
100 Chapter 4 Transducers and Control System Components

magnetic pick-up device that senses the passage of an individual paddle as it


sweeps past the pick-up. The pick-up device is comprised of two components.
One is a magnet whose magnetic field becomes distorted by any nearby metallic
movement (a paddle). In the process of becoming distorted, this field sweeps past
a second component, a pick-up coil, causing a voltage to be induced inside the
coil. Therefore, for every passage of a paddle wheel past the pick-up, there is a
voltage pulse that is generated by the pick-up itself. Then it becomes a simple
matter of correlating the number of pulses per second to the flowrate of the
passing fluid in the pipe or tube in which the transducer is installed.
Another common method for the sensing of gases uses what is referred to as
the hot-wire anemometer principle. In this process (Figure 4-16) a heated wire is
allowed to be cooled by the flowing gas. The amount of cooling is determined by
the amount of decrease in the wire’s electrical resistance. (Remember that most
pure substances decrease their electrical resistance when cooled.) Once this
change in resistance is known, this can then be correlated to the flowrate of the
gas flowing past the sensor itself.

Another more indirect method for measuring the flowrates of fluids uses a
created pressure drop for its flowrate detection. As fluids flow past pressure taps
installed on both sides of an orifice in a pipe, a pressure difference is created
between the two taps as a result of this orifice. A manometer device is attached to
both pressure taps resulting in the pressure drop that exists across the orifice to be
sensed by the manometer. The manometer usually contains some sort of electri¬
cal sensing device, such as an LVDT, to detect the height or pressure head of the
manometer’s reading fluid. Figure 4-17 shows such a system.
Section 4-7 Electromagnetic Sensing 101

LVDT

Flowrate
Indicator

Float

Figure 4-17 Using a pressure differential for determining liquid flowrates.

The transfer functions for all of the aforementioned flow-rate-sensing units


are the same. Namely,

£out _ ^flowrate
(4-8)
Q 1 + ST

where -^fflowrate sensor constant or transfer function value for flowrate sensor.
Must be obtained from manufacturer or obtained experimen¬
tally.
Q fluid flowrate expressed either as weight flow rate (lbs/sec) or
volumetric flow rate (ft3/sec)
s Laplace transform for jo
T time constant of sensor (sec)

4-7 ELECTROMAGNETIC SENSING

Electromagnetic sensing covers a very large range of sensor types. This is


due to the very broad nature of the electromagnetic spectrum. This is the broad
illustrated band of radio and light frequency designations, called the electromag¬
netic; spectrum, often seen illustrated in physics textbooks and illustrated in Fig¬
ure 4-18. But we limit our electromagnetic sensing to just the radio communica¬
tions frequencies, since the other portions of this spectrum already have their own
unique sensor types. For example, look at the visible light portion of the spec¬
trum. There are special light-sensitive sensors that have the task of responding
ir
h— >
o o
z 77
QJ C/3
: z
UJ —
_j oc N
LU UJ n h_
> S a:
< w £uj
51 ll i.

<
5 !® CN
5< ”6
<[ x
a o
*,
-o

if)
>- o

1X o
yc X
-o

*—
UJ
—M
o
> o
<
IX X
j -©
3

iHon
319ISIA «D
—O

-o

Q
IULU
X
< ■'T
X —o
IUL o.
z -o

CO
im..
> CN
<
§ o
o
X —2
o

—6
UJ >^ X
> !UL -o
<

O ^ UM
o o>
< X $
-I CO ^
—o

< — 1

O uu
z> —o
el -o

o
__s
x »- z
UU IUU < uJU N
?zz: -o

<

102
Section 4-7 Electromagnetic Sensing 103

only to visible, or nearly visible, light radiation. And in the X-ray portion of the
spectrum, there are sensors that are devoted solely to the detection of X-rays. We
concentrate only on the radio frequency sensing devices for this particular dis¬
cussion.
Basically, a radio frequency sensing device (or rf sensing, as it is usually
called) is nothing more than a radio receiver that has been designed to tune a
desired range of radio frequencies. There are a wide variety of radio receiver
circuits available, and many have been constructed out of a single integrated
circuit similar to the one depicted in Figure 4-19. In general, an rf receiver's
circuitry can be subdivided into the following individual circuits: a radio fre¬
quency amplifier section, which is usually tunable for a range of frequencies; a
detector section that “strips” the audio off the carrier: and an audio amplification
section. The only input adjustments needed to the IC are for tuning and for
adjusting the receiver's sensitivity. In many cases, the receiver is purposely de¬
signed to be broadbanded or tuned so that no tuning is necessary at all. Wide
ranges of frequencies can then be received simultaneously without having to tune,
as long as it isn’t necessary to receive information from more than one frequency
carrier at a time that might be transmitted simultaneously on any of the many
received frequencies.
Typical applications requiring the use of rf sensing in an automatic control
system would be the automatic tracking of a moving radio-wave-emitting object
such as an airplane, missile, or satellite. These are fairly sophisticated systems
and are not discussed in this book.
In order to develop the transfer function for an rf receiver, we must first look
at the description for the incoming radio signal to the receiver. The radio signal's
strength or quantity is usually given in units of the microvolt. The output of the
receiver is usually given in volts. Consequently, the receiver's transfer function
would then be:

°UtpUt-V-f-S = K, (4-9)
input volts
where K,f = gain or transfer function value for an rf receiver (voltslyevolt)
output volts = output voltage of receiver (DCV)
input volts = input signal strength (ACq.V)

Figure 4-19 A typical IC package (a


DIP or dual in-line pin configuration)
for housing an entire FM radio re¬
ceiver.
104 Chapter 4 Transducers and Control System Components

4-8 TEMPERATURE SENSING

There are a number of methods available for converting temperature into an


electrical signal. One method is through the application of a solid state device
called a thermistor. Most pure elements in nature increase their electrical resis¬
tance when exposed to an increase in temperature, with carbon being the most
notable exception. Many compounds, however, react oppositely. And in particu¬
lar, the compounds referred to as semiconductors, which are used in the produc¬
tion of semiconductor devices, react with this opposite behavior. That is, their
electrical resistances decrease with an increase in temperature.
The thermistor is one such device (Figure 4-20). Thermistors are relatively
inexpensive and can be placed into very small recesses for temperature measur¬
ing. A typical calibration curve correlating a thermistor’s DC resistance to tem-

Figure 4-20 Thermistors used for


sensing temperature. (Courtesy of
Omega Engineering, Inc., Stamford,
Conn.)

perature would look like the ones in Figure 4-21. It would be rare, however, to
use a thermistor’s resistance measurement for a direct temperature reading.
Instead, a current flow or a voltage drop (most likely a voltage drop across a series
resistance) would be used for a temperature indication. A thermistor’s transfer
function would then be determined in the following manner:

^thrm
(4-10)
°F 1 + ST

where Kthrm sensor constant or transfer function of thermistor. Must be ob¬


tained from manufacturer or obtained experimentally.
AC or DC voltage output
the temperature amount being sensed (°F)
s Laplace transform for jco
T time constant (sec)

Another commonly used temperature sensing device is the thermocouple.


This device depends on a phenomenon called the Seebeck Effect. This is where
one can take two dissimilar metal wires and, by heating the ends of these wires
where they have been physically joined either through twisting or welding, pro¬
duce an electromotive force (emf) at the wire’s opposite ends. The emf, or volt¬
age, being generated is quite proportional to the amount of heat being applied to
the junction. Since this voltage is quite small, in the millivolt range typically, this
voltage is usually amplified to a larger value and represented either by a larger DC
Section 4-8 Temperature Sensing 105

Part No. 44004 Part No. 44005 & 44007

o CJ
+1
Q.
TEMP

HI

TEMPTC TEMPTC

Part No. 44006

O
TEMP. • "C

Q.
LLI

-80 40 0 40 80 120 150


TEMPTC TEMPTC
”C
TEMP

-60 -40-20 0 20 40 60 80 100 -60 -40 -20 0 20 40 60 80 100


TEMPTC TEMPTC
"C
TEMP

-60 -40 -20 0 20 40 60 80 100 -60 -40 -20 0 20 40 60 80 100


TEMPTC _RESISTANCE ±% TEMPTC
_TEMPERATURE ±°C

Figure 4-21 Response curves for thermistors. (Courtesy of Omega Engineering,


Inc., Stamford, Conn.)
106 Chapter 4 Transducers and Control System Components

or AC voltage. Consequently, the transfer function units of the thermocouple


would be identical to the thermistor’s transfer function. Namely,

A\cpi
(4-11)
1 + ST

where A'tcpi = sensor constant or transfer function value. Must be obtained from
manufacturer or obtained experimentally.
= DC voltage output (mV)
= the temperature amount being sensed (°F)
5 = Laplace transform for jcu
T = time constant (sec)

There are other ways of detecting temperature, some of which are rather
sophisticated, but we consider only the two aforementioned methods. These
cover a fair percentage of all temperature detection systems.

4-9 LIGHT SENSING

In the case of light sensing, there are several means available for the detection of
its intensity. As a result, light sensing may be broken down into roughly three
different categories:

1. the photoresistive sensor,


2. the photoemissive sensor, and
3. the photovoltaic sensor.

We discuss each without getting too deeply involved with the physics or chemis¬
try of how each works. However, before we do this, we first define the unit of
light intensity, since this is going to be the input quantity to all of our transducers
that we discuss in this section.
The unit of illumination can be quite confusing and is not too well under¬
stood by some people. For the purpose of our discussion, we define the unit of
illumination as the flowrate of light. In this context, think of light as being com¬
prised of lines of flux, just as you thought of magnetic lines of force as flux when
dealing with the discussion of magnetic fields in electricity. The input of this light
flowrate is the lumen, a unit of measurement in the Meter-Kilogram-Second
(MKS) system of units. Having now defined the unit of illumination, let’s con¬
tinue now with our discussion of light sensing transducers.

4-9.1 The Photoresistive Sensor

This photoresistive device (Figure 4-22) varies its internal DC resistance accord¬
ing to the light intensity falling on the sensor. Its chemistry is somewhat similar to
Section 4-9 Light Sensing 107

Figure 4-22 The photoresistive sen¬


sor. (Courtesy of Motorola Semicon¬
ductor Products, Inc., Phoenix, Ariz.)

that of the thermistor except that the wavelengths of electromagnetic energy


response are shorter. (Refer back to the electromagnetic spectrum chart in Figure
4-18.) Like the thermistor, however, the change in resistance response is usually
transformed to a varying voltage for the correlation to light intensity.
As a result, the sensing transducer containing the photoresistor has the
following transfer function:

K,photo
(4-12)
Q 1 + ST

where Kphoto constant or transfer function value for photoresistor transducer.


Must be obtained from manufacturer or obtained experimentally,
the input flowrate of light (lumens)
the output voltage of the transducer in either AC or DC volts
5 Laplace transform for jco
r time constant (sec)

The substance the light sensitive material is made from will determine that
sensor’s time constant, r. Photodiodes and phototransistors have time constants
108 Chapter 4 Transducers and Control System Components

less than 1 ps. Lead selenide sensors have constants in the 10 ps range, whereas
lead sulfide cells are in the 100 to 1,000 ps range. Cadmium selenide cells are
typically as high as 10 ps, whereas cadmium sulfide cells typically run as high as
100 ps.

4-9.2 The Photoemissive Sensor

The photoemissive transducer uses a light sensing device that emits electrons in
some proportion to the incoming flowrate of light. These transducers have the
unique characteristic of being extremely sensitive to weak levels of light. Other
than that however, the transfer function is identical to that of the photoresistive
sensor.

4-9.3 The Photovoltaic Sensor

The photovoltaic transducer uses a light sensing material that has the capability of
producing an emf that is proportional to the incoming flowrate of light. A unique
characteristic of this material is that the generated voltages are extremely linear to
the lightflow amounts. However, as in the case of the photoemissive sensor, the
transfer function is identical to that of the photoresistive sensor. Typical time
constants are 20 ps for this type of cell.

4-10 OTHER CONTROL SYSTEM COMPONENTS:


SERVOMECHANISM DEVICES

Up to this point, we have been concentrating our discussions on sensing devices


and their transfer functions. Now, we consider a group of devices that are used
extensively in automatic control systems, namely, the servomechanism. A servo¬
mechanism is an electromechanical system comprised of devices designed specifi¬
cally for accurate positioning or the controlling of motion. These devices are
manufactured to close mechanical and electrical tolerances and may be consid¬
ered to be the very heart of the automatic control system.

4-11 THE SERVOMOTOR

The first servomechanism device that we study here is the servomotor. The ser¬
vomotor may be thought of as a precision electric motor whose function is to
cause motion in the form of rotation or linear motion in proportion to a supplied
electrical command signal. There are two general types, the AC servo and the DC
servo. DC servos are used generally for high power applications requiring large
amounts of torque capability. However, in addition, DC servos are noted for
generating radio frequency (rf) interference and requiring a certain amount of
Section 4-11 The Servomotor 109

maintenance, such as the periodic changing of their brushes. And since these
motors require DC for their operation, the DC amplifiers needed to run these
motors have had a tendency in the past to drift, thereby suffering from a reliability
problem. Recent developments in DC amplifier design, however, have produced
highly stable amplifiers. AC servos, on the other hand, tend to be more stable in
their operation. They also tend to be lighter in weight, somewhat more rugged,
and require less maintenance. However, they do lack the large torque capabilities
of the DC servo.
Figure 4-23 shows a view of a typical servomotor used in small automatic
control systems. We now derive the transfer function for the AC servomotor and
try to gain some insight into how this motor actually works.

Figure 4-23 Servomotor. (Courtesy of


The Singer Company, Kearfott Guid¬
ance and Navigation Division, Little
Falls, N.J.)

To begin with, the AC servomotor is considered to be a two-phase induction-


type motor. It’s an induction motor because the rotor is inductively coupled
(similar to the coupling that exists between the windings of a transformer) to the
stator rather than through brushes (as in the case of a DC motor), and it’s a two-
phase motor because of the two separate voltage source windings that it possesses
(Figure 4-24). In reality, the servo has more than just the two windings. Typi¬
cally, four windings are used. Figure 4-25 shows how they are all wired together.
The servo’s speed of rotation is determined by the amplitude of the incoming
control signal being supplied from the servo’s amplifier. Its speed can also be
dependent on the control signal’s frequency. An amplifier is usually needed,
preceding the motor, to supply the control signal since a sizeable voltage and
current signal is needed to produce the power to drive the servo. It is rare to find a
control signal being supplied directly from a transducer or other nonamplifying
device having this necessary power capability.
The servo’s direction of rotation is determined by the phase relationship
existing between the control signal current and the power line current signal, in
this case, called the fixed phase current. Figure 4-26 shows the waveforms of the
110 Chapter 4 Transducers and Control System Components

■O

Fixed Phase Fixed AC Voltage


Winding Supply

o
-'THHP—
Control
Phase
Winding

0 6
Variable Signal
from AC Amplifier

Figure 4-24 Schematic of an AC servomotor. (Photo courtesy of The Singer Company,


Kearfott Guidance and Navigation Division, Little Falls, N.J.)

two currents being supplied to the servo’s windings. Notice that the control phase
current leads the fixed phase current by 90°. This particular relationship is normal
and is created by the servo amplifier that is feeding the control phase signals to the
servomotor. This also causes the servomotor’s output shaft to rotate in a particu¬
lar direction. Now, if the control phase current is shifted 180°, the resultant
waveforms would look like Figure 4-27. The control phase current signal will now
lag the fixed phase current by 90°. This new relationship will cause the servomo¬
tor’s output shaft to rotate in the opposite direction from the previous direction.
The reason for this reversal in rotation has to do with the revolving fields created
by the changing stator currents and the 90° control phase differences within the
servomotor. A discussion of this is beyond the purpose of this text but may be
found in any good small induction motor theory text.
Section 4-11 The Servomotor 111

Fixed
Phase

Figure 4-25 How the four windings of


Phase an AC servomotor are connected.
Current

Figure 4-26 The control phase current


leads the fixed phase current by 90°.

Figure 4-27 The control phase current


lags the fixed phase current by 90°.

4-11.1 Calculating the Transfer Function for a Servomotor

In order to develop the transfer function for a servomotor, we must first under¬
stand the concept of torque. Torque is defined as the product resulting from
multiplying a force being applied at a right angle to a moment arm by the length of
112 Chapter 4 Transducers and Control System Components

that moment arm. For a motor, its output torque for any given speed of shaft
rotation would be equal to the force that would be produced at the end of an arm
attached to the motor’s shaft, multiplied by the length of that arm. This is the
length, /, indicated in Figure 4-28.

The torque curve for a motor is the curve relating that motor’s torque output
with the speeds at which the torque measurements were made. A typical curve
for an ordinary induction motor is seen in Figure 4-29. Also shown is the torque
curve for an induction servomotor. The stator winding’s DC resistance is typi¬
cally several times more than the stator of an ordinary induction motor. The
effect of this is to cause the shape of the servomotor’s torque curve to straighten
out considerably, as can be seen in Figure 4-29. This will make mathematical
approximations somewhat easier for us when we have to use this information later
on, because we can approximate these curves with straight lines. In fact, many of

Speed

Figure 4-29 Torque curves for an


ordinary induction motor and an AC
servomotor.
Section 4-11 The Servomotor 113

the torque curves seen in catalogs published by servomotor manufacturers are


already straight lines. Figure 4-30 is a typical example.
Let’s take a further look at Figure 4-30. Notice that there are several speed-
torque curves, each curve representing the speed-torque for a particular control
voltage value. In reality, servomotors operate at fairly low voltage values. In
other words, the signal voltages supplied to the servo are usually quite small,
certainly less than 25% of the servo’s rated control voltage. The reason for this is

Figure 4-30 Typical servomotor


straight-line speed-torque curves found
Torque (oz.-in.) in many manufacturers’ catalogs.

because most of any automatic control system’s time is spent making minor
touch-up adjustments to maintain a particular position, pressure, temperature,
etc. In the case of the systems we study here, these small adjustment voltages are
continuously and automatically being fed to our servomotor. These voltages will
be found to be switching back and forth in polarity. So what the servomotor
experiences are small nudging voltages causing the servo to jockey back and
forth, performing what closely resembles a balancing act, in order to maintain a
certain desired shaft output location or position.
Looking at the 25% curve in Figure 4-30, note where this straight line meets
the 0 rpm level at approximately the 0.8 torque value. This torque value is called
114 Chapter 4 Transducers and Control System Components

the motor’s stall torque. If you were to supply this motor with 25% of 26 VAC for
its control signal voltage, and hold down its shaft rigidly, you would experience its
stall torque for that particular control voltage.
Looking at Figure 4-30 once again, notice the intersection of the upper
portion of each curve with the left-hand vertical axis. This intersection point, as
read on the speed axis, is the free-wheeling speed (i.e., the torque is zero) for that
particular control voltage.
Now that we have analyzed a servomotor’s speed-torque curve, its free¬
wheeling speed, and its stall torque value, we now must define yet another impor¬
tant characteristic before deriving our transfer function. This characteristic is
called the motor’s viscous damping. Viscous damping is a resistance that the
servomotor’s shaft experiences when it rotates. This resistance is created by an
interaction that takes place between the magnetic fields of the rotor and stator.
(Viscous damping can be observed by applying full fixed phase voltage to the
proper set of windings, and shorting out the control phase windings. Next, try
rotating the motor’s shaft. The resistance felt is the servomotor’s viscous damp¬
ing. The faster you try rotating the shaft, the greater the resistance that you feel
becomes.)
Viscous damping could be considered good or bad depending on your view¬
point and application. It may be considered bad from the standpoint that it robs
the motor’s output of power; it may be considered good because it reduces over¬
shoot, a condition that we discuss in a later chapter. For now, though, we want to
be able to calculate the magnitude of this damping effect, because it has a pro¬
found effect on the motor’s transfer function.
The amount of viscous damping, Dv, is calculated by the following equation:

(4-13)
OJf

where Dv = viscous damping (oz-in/rev/min or dyne-cm/rad/sec)


Ls = stall torque (torque at c* = 0) at 25% rated control voltage (oz-in or
dyne-cm)
ajf = free-wheeling speed (torque = 0) at 25% rated control voltage (rad/
sec)

Note: oz-in/rev/min = ounce-inch per revolution per minute,


dyne-cm/rad/sec = dyne-centimeter per radian per second.

Notice that both English and Centimeter-Gram-Second (CGS) units have


been listed for the terms in Eq. (4-13). To convert oz-in/rev/min to dyne-cm/rad/
sec, multiply by the conversion factor, 6.742 x 105.
We now must calculate the amount of torque lost due to the viscous damp¬
ing. This is done by letting:

Lv coDv (4-14)
Section 4-11 The Servomotor 115

where Lv = torque lost due to viscous damping (oz-in or dyne-cm)


to = rotational velocity of shaft (rev/min or rad/sec)
Dv = viscous damping from Eq. (4-12) (oz-in/rev/min or dyne-cm/rad/sec)

Now, we must define the servomotor’s torque constant, KLM. This is done
with the following equation:

Klm = ~~~ (4-15)


t-cv

where KLM = the motor’s torque constant (oz-in/volt or dyne-cm/volt)


Ls = the motor’s stall torque at 100% control voltage (oz-in or dyne-cm)
ecv = 100% control voltage (volts)

In general, therefore:
^produced Ls Lv

Or ^produced (^Lm) " (^cu) (oV (4-16)


Now the produced torque, Lproduced, must equal the absorbed torque, ./-absorbed-
The absorbed torque is comprised of the inertial torque of the rotor, Ja, and the
bearing friction torque (which is negligible, unless the bearings are gone). We are
also assuming that the load connected to the motor’s shaft is zero. This is an
assumption that is not too far from the truth. Most servomotors work through a
gear transmission, and because of this the motor “sees” very little torque load.
We can now write:
(-^Z-mX^cu) tx)Dv JoL

or (KLM)(ecv) = Jot + mDv (4-17)


where J = moment of inertia of motor rotor (slugs-ft2 or gm-cm2)
a = angular acceleration (rad/sec2)

Before going on, let’s identify what our goal is so that we know what it is that
we want. To begin with, we want a transfer function that says:
output e
- K SM (4-18)
input cv

where 0 = angular displacement of shaft (deg)


ecv = control phase voltage (volts)
Ksm = transfer function value (deg/volt or rad/volt)

We now resort to Laplace transforms to write the final equation for our transfer
function. Referring to Table 3.1, we find the equivalent Laplace transform ex¬
pressions for the following terms:
a = s26
oj = sO

€CV ECV
116 Chapter 4 Transducers and Control System Components

Therefore, KLMECV = Js26 + DVS

or KlmEcv = 6(Js2 + Dv)

Solving for 6/Ecv:

0 _ klm
Ecv Dvs + Js2

There is some additional cleaning up that we can do to make the final trans¬
fer function of a servomotor appear in a more standard form. Let r be the motor’s
time constant. This is the time it takes for the motor to reach 63.2% of its intended
speed after receiving 100% of rated control voltage signal. Then,

T = jr (4-19)
tO v

where r = servomotor’s time constant based on 100% rated control voltage signal
(sec)

We also define now a new term, velocity constant, Kv:

= (4-20)

or

Klm = K v

therefore

_6_ = Ky • Dv

Ecv Dvs + Js2

Note: The velocity constant may also be calculated using the following equation:

Kv = y1- (4-21)

This equation is not used in this particular transfer function development, but is
used later.
Now, multiplying the 6/Ecv right-hand expression by 1IDV, (both the numera¬
tor and the denominator), we get:

j:
Js2
5 +

Because JIDV = r (Eq. 4-19),

_0_ = Ky

Ecv S + TS2
Section 4-11 The Servomotor 117

Ky rad
Finally, Kvdsm — (4-22)
s(l + TS) volt

where KVDSm = transfer function value for viscous damped servomotor (rad/volt)

Equation 4-22 represents the transfer function for a servomotor assuming


that you know its time constant, r, and its velocity constant, Kv. These values
can be obtained from the manufacturer’s catalog data, except for one problem.
Remember when we based all of our operating conditions for a servomotor on its
25% of rated control voltage line rather than using the 100% line as the manufac¬
turer does? Unfortunately, much of the published data in a catalog is based on
this 100% line. We have to make adjustments to the data to place it into the more
realistic 25% of rated control voltage operating area. To do this, we make the
following adjustments to any catalog data that we see:

Actual Dv = \dv\ (4-23)

Actual t = 2t ! (4-24)

Actual Kvm = 2Kvm ! (4-25)

Actual J = J! (4-26)

The exclamation mark (!) indicates that the information is unconverted catalog
data. We use this convention from now on.
Bear in mind that the time constant that we are talking about here is a
mechanical time constant, not an electrical time constant. Both, however, have
similar characteristics. That is, one time constant equals 63.2% of the full reac¬
tion time. It takes approximately 5t to equal the full reaction time. For a com¬
plete transfer function that takes into account both the mechanical and electrical
time constants, the function would look like this:

6 _ Kv rad
(4-27)
Ecv s(l + ts)(1 + TeS) volt

where re = servomotor’s electrical time constant (sec)

EXAMPLE 4-2 ---


Derive the transfer function for a servomotor given the following catalog informa¬
tion: r! = 0.0147 sec., the no-load speed (ay!) is 4,900 rpm, and the rated control
voltage (Ecv\) is 36 VDC.

Solution:
First, calculate Kv\. Do this by first converting the 4,900 rpm figure to rad/sec so
that we can use Eq. (4-21):
118 Chapter 4 Transducers and Control System Components

(4,900)(2tt)
(Of =-^-rad/sec

= 513 rad/sec

, 513 rad/sec
Then, using Eq. (4-21):
Kv-~~36~^\T
= 14.25 rad/sec/volt

And since

Kv = 2Kv \ (Eq. 4-25)

then Kv = 2(14.25) rad/sec/volt


= 28.5 rad/sec/volt

Next, we have to convert the time constant for the servomotor into actual data. We
do this by using

t — 2t ! (Eq. 4-21)

Therefore, r = 2(0.0147) sec

= 0.0294 sec

Finally, using Eq. (4-22):

Q 28.5
Ecv ~ 5(1 + 0.0295)

The type of servomotor that we have been describing here uses viscous
damping that is applied continuously throughout its operation. This means that
the motor has to work continuously against this magnetic damping whether it
wants to experience damping or not. There is another form of this motor which
has viscous damping applied only when there is a change in speed. From the
standpoint of efficiency, this is certainly more desirable. This type of servomotor
is called an inertially damped servomotor. Figure 4-31 shows a diagram of this
device. It works like this: Surrounding the motor’s output shaft is a series of
magnets that is free to rotate on a circular raceway. The magnetic field from this
arrangement interacts with a similar set of magnets rigidly mounted on the shaft
itself. As long as the shaft rotates at a constant speed, the outer raceway magnets
will rotate at that same speed. Under this condition, few, if any, magnetic lines of
force are being cut. However, should the shaft’s speed suddenly change, the
magnetic fields become cut, causing an opposing viscous damping torque to be
generated. This in turn causes the shaft’s rotation to slow down.
The transfer function for this type of servomotor becomes somewhat more
complicated. There is more than just one time constant that becomes envolved
here. The transfer function now becomes:
0 out Kyi 1 + st2) rad
K IDSM (4-28)
'CV 5(1 + ST])(1 + 5T3) volt
Field Freewheeling

Figure 4-31An inertially damped servomotor (IDSM). (Photo courtesy of The


Singer Company, Kearfott Guidance and Navigation Division, Little Falls, N.J.)
120 Chapter 4 Transducers and Control System Components

where n, r2, and r3 = mechanical time constants for the servomotor (secs.)
Kidsm = transfer function value for inertially damped servomotor
(rad/volt)

4-12 THE GEAR TRAIN

It perhaps may seem a little strange to introduce the subject of gear trains at this
point, since gear trains appear to have nothing to do with a discussion on precision
electromechanical devices. However, precision gear box transmissions play an
extremely important role in the design of automatic control systems. Rarely do
you hook up or couple a servomotor directly to some mechanical device to be
driven by the servo directly. The servo’s shaft speed would be entirely too high.
Typically, these shaft speeds can exceed 10,000 rpm. For a motor to operate at
that speed and perform precision responses to command signals would be virtu¬
ally impossible. Consequently, servomotors generally operate through geared-
down transmissions, or gear trains, whose purposes are to furnish the more desir¬
able slower shaft speeds and to create an increase in needed output torque.
Whenever you purchase a servomotor, you will most likely be asked to specify
your preference in gear ratios to be supplied with the servo. The geared transmis¬
sions can be easily attached or exchanged on these servomotors to satisfy the
controls hardware design when needed.
Let’s look more closely at a gear train. In Figure 4-32 we see several gears
that are in mesh forming a simple gear train mechanism. Gear 1 is the input gear
and gear 4 is the output gear. Gears 2 and 3 couple the rotation of gear 1 to gear 4
but really don’t enter into the transfer function that we develop here. Again, using
the relationship that a mechanism’s transfer function equals input/output, and
because we are interested in the displacement, 6, that takes place between the
gear train’s input and output, the transfer function is going to depend on the gear

Gear 4
Output
Rotation

Gear 3

Input
Gear 1 Rotation

Figure 4-32 A gear train.


Section 4-12 The Gear Train 121

train’s gearing ratio. The gear ratio, in turn, is going to be dependent on the gear
tooth numbers existing between the driver gear and driven gear in our gear train.
In other words:

output2 02 _ N±

input] 6\ N2

output3 03 _ A/2
Also,
input2 62 A/3
output4 64 __ A/3
Also,
input3 62 A/4
Then, the overall output/input transfer function will be equal to the products of
the individual output/input functions. Or,

output output2 output3 output4


input input] input2 input3

output output4 04
Or,
input input] 01

Then, 04
01

0QUtpUt
In general, for any gear train: (4-29)
0jnput

Equation (4-29) tells us that regardless of the number of gears existing inside a
gear box or transmission, its transfer function is dependent only on the ratio of its
output gear displacement versus its input gear displacement. We have to be care¬
ful here, though. We don’t want to confuse a gear train’s transfer function with its
gear ratio. The gear ratio, sometimes called the speed ratio of a gear train, is
equal to its input gear teeth number divided by its output gear teeth number. In
many instances, this ratio is expressed as the letter N. Also, N represents the
ratio of the output speed divided by the input speed; this ratio, as it turns out, is
the reciprocal of the gear train’s transfer function. Summarizing,

ts _ 0QUt T\in ^out J_


J^gear ~ a (4-30)
v\'in T,out CO in N

where Kgear the transfer function value for a gear train (no units)
0OUt output gear’s displacement (rad)
0in input gear’s displacement (rad)
1 ^out output gear’s angular velocity (rad/sec)
COin input gear’s angular velocity (rad/sec)
rout output gear’s tooth number
Tin input gear’s tooth number
N gear train’s speed ratio or tooth ratio
122 Chapter 4 Transducers and Control System Components

Another often used gearing system in automatic controls applications is the


rack and pinion. Its construction is shown in Figure 4-33. The rack and pinion is
used to change the direction of motion and to change rotary motion or displace¬
ment to linear motion or displacement. Its transfer function is:

K _ output _ linear displacement inches (4 31)


rp input rotary displacement rad

where Krp = transfer function value of rack and pinion combination (in/rad)

Figure 4-33 The rack-and-pinion gear


Movement of Rack drive system.

4-13 RATE GENERATORS

A rate generator is nothing more than a small precision-made electrical generator


whose function is to produce an electrical voltage in proportion to its rpm. Other
common names for this device are the tachometer, servo-generator, and rate
generator. There are three basic applications for this device:

1. as a speed controller device,


2. as a stabilizing device for position control, and
3. as a mathematical integrator.

The most common application for this device is probably as a speed controller.
This application requires the generator to control and maintain a particular rota¬
tional velocity of a mechanism. The output voltage of the rate generator is fed
back to the mechanism in such a way to nullify or cancel a portion of that mecha¬
nism’s supply voltage. Any increase in supply voltage would then be cancelled,
thereby serving to hold the same velocity. This, if you will recall, is an application
of negative feedback, which we discuss in detail in a later chapter.
Rate generators are manufactured for the purpose of generating both DC and
AC voltages. The AC generator is the more commonly used of the two types.
Figure 4-34 shows a simple wiring layout for an AC generator. Because of its
Section 4-14 Synchromechanisms 123

Excitation or
Fixed-phase Coil

? 6fp ?

-dlfiJL'-

Control Phase or
e°p Output Coil

Figure 4-34 An AC two-phase rate


-o generator.

frequent use as a speed controller and as a positional controller, the rate generator
is often built as an integral part of a servomotor.

4-13.1 The Rate Generator’s Transfer Function

The transfer function value for the rate generator, Kg, is determined in the follow¬
ing way:

Kg = — (4-32)
Ct)jn

where Kg = transfer function value for rate generator (volts/rad/sec or volts-sec/


rad)
Eout = output voltage (volts)
coin = input rotational velocity (rad/sec)

4-14 SYNCHROMECHANISMS

This group of devices is somewhat off into its own area of special automatic
control devices. This is a hardware group comprised of those electromechanical
devices that have the capability to transduce position into an electrical signal. In
these devices are found a set of phasing coils for a stator, a supply voltage coil for
a rotor, and an output shaft. Synchromechanisms usually operate in pairs, one
unit referred to as the transmitter while the other is called the receiver. Their
theory of operation can be somewhat complicated, so we discuss the theory only
superficially here.
To begin with, synchros can be divided into two general groups depending
on their application. These groups are:
i

1. control synchros, for indicating readings of position or location from a re¬


mote location; and
2. torque synchros, for performing work based on remotely transmitted sig¬
nals. Torque synchros usually use servomotors for their operation.
124 Chapter 4 Transducers and Control System Components

The only difference between these two groups is in their internal construction.
The torque synchro uses heavier wire in its windings and is generally more rug¬
gedly built. It is used for the transmission of power. The control synchro is
usually more lightly constructed and, consequently, smaller in physical size be¬
cause it doesn’t require the carrying of much current in its windings.
Figure 4-35 shows the schematic of a synchro. Note that there is a control
transmitter (CX) and a control receiver (CR). (The discussion here would be the
same for a torque transmitter, TX, and a torque receiver, TR). Basically, the
theory of operation for this system is based on the action of a transformer. It is
important to realize when considering Figure 4-35 that the stator voltage’s magni¬
tude and phase are determined by the angle of rotation of the rotor relative to the
stator. It is this relationship that permits the CX (or TX) to accurately transmit its
rotor position to the matching CR (or TR). The receiver’s rotor will respond to
the voltages transmitted to its three windings and will automatically rotate, due to
motor action. It will continue to do this until it finds a spot in its rotation where
the three transmitted voltages and the supply voltage produce a torque that is
essentially zero. This will cause the CR synchro’s shaft to come to rest at pre¬
cisely the same location as the CX’s shaft.

Supply Voltage

Figure 4-35 How the synchromechanism operates.

Figure 4-36 shows a typical application of a remote reading synchromech¬


anism. The CX is attached to the dial mechanism of a Bourdon-tube pressure
gage. The CR is located at a remote data site and is attached to a dial that would
have normally been installed in the pressure gage. As the pressure gage responds
to some pressure, the CX transmits the amount of rotation of the Bourdon mecha¬
nism to the CR where the actual pressure reading can then be read.
Section 4-14 Synchromechanisms 125

Three-phase Wiring

Pressure
Input

Figure 4-36 A remote reading pressure gage using a CX-CR synchromechanism


pair.

4-14.1 The Transfer Function for Synchros

Because the input to a CX or TX is rotational motion and the output of these two
devices is voltage, the transfer function for an individual CX or TX synchro
becomes:

E0 ut ^ volts
(4-33)
1 ~ ~ Kcx,tx

where KCxnx = transfer function value for either a CX or a TX synchro (volts/


rad)
0m = input displacement of rotor shaft (rad)
E0ut = output voltage of synchro (volts)

The transfer function for a CR or TR synchro becomes the inverted equivalent of


the units’ transfer function described in (Eq. 4-33), since the input now becomes a
voltage and the output becomes a rotational motion. Namely:

,
0 out rad
(4-34)
— Kcr/tr
'in volts

It’s a little unusual to see a CX or CR synchro unit being used by itself. And
it’s just as unusual to see a CR or TR synchro being used by itself. Normally,
these units are used in pairs, since one unit depends on the other units to complete
a remote reading system. However, their individual transfer functions are given
126 Chapter 4 Transducers and Control System Components

here as more of an exercise than anything else. Of course, what has been said of
the CX-CR systems is also true for the TX-TR systems.

4-14.2 The Synchro Control Transformer

A synchro control transformer, or CT, is electrically quite similar to the synchro


control transmitter (CX) and receiver (CR). However, mechanically it operates
quite differently. It has an input shaft and winding construction much like the CX,
to indicate position. However, on its other shaft is a winding whose output is an
AC voltage that is proportional to the difference between the rotational displace¬
ments of the input and output shafts (see Figure 4-37). It may help to think of the
CT synchro as being similar to a CX-CR synchro all in one unit with the common
rotor connections removed. The CX rotor winding, instead, becomes the input

Shaft Shaft

Figure 4-37 Using a CT to generate an error voltage. (Photo courtesy of The


Singer Company, Kearfott Guidance and Navigation Division, Little Falls, N.J.)
Section 4-16 Auxiliary Servo Circuits 127

connections for the CT where an input AC voltage is applied, while the CR


windings are the output connections. The output AC voltage is then proportional
to the relative displacements of the two input shafts, the CX rotor shaft and the
CR rotor shaft. This output voltage is called an error voltage and is frequently
used in automatic control systems for controlling position.
The transfer function for a CT system, KCt, is somewhat similar to the CX or
TX system. The only difference is that the rotational input for the synchro is an
expression of a rotational difference existing between its two input shafts.

4-15 RESOLVERS

Resolvers are another group of synchros that can be classified under the heading
of synchromechanisms. They are used for the purpose of performing trigonomet¬
ric calculations and may be thought of as analog electromechanical computers.
They can transform coordinates and perform conversions between polar and rec¬
tangular coordinates.
A resolver is like a variable transformer, similar in construction to that of the
synchros just discussed. However, unlike the CX-CR systems requiring a three-
phase coil system for a stator, the resolver can have as few as two coils for its
stator winding and the same for its rotor winding. The coil pairs are wound at
right angles to each other, as seen in Figure 4-38.
Resolvers, like the CX-CR or TX-TR systems, operate in pairs. However,
generally no distinction is made between which unit is a transmitter and which is
the receiver. Looking at Figure 4-38, we see how a typical resolver system might
be wired. In this case, two resolvers are being used to determine the sine and
cosine of two angles, a and (3. The system would operate like this:

1. A fixed supply voltage would be applied at Es i_2.


2. Rotor 1 would be rotated through an angle, a.
3. Rotor 2 would be rotated through an angle, /3.
.
4 The voltage resulting at ER t.2 would then be = Es i.2cos(a + (3).
5. The voltage resulting at ER 3.4 would then be = Es i-2sin(a: + f3).

If the wires were crossed over at the two spots marked with Xs in Figure
4-38, the two resolvers would then produce entirely different results. The resul¬
tant voltage at ER j_2 would then become Es i.2cos(o: - f3); the resultant voltage at
Er 3.4 would become Es i.2sin(a - (3).

4-16 AUXILIARY SERVO CIRCUITS

Up to this point, we have discussed the major servomechanical components used


in automatic control systems. However, there are several electronic circuits that
are frequently used to modify or to create certain characteristics within portions
128 Chapter 4 Transducers and Control System Components

Er 1-2 ^S1-2 C0S ER2-4 ^S1-2 s*n $)

Jumper-short

Figure 4-38 A resolver system used for the addition of two angles.

of the control system, just as there are electromechanical devices used for the
same purpose. In the next few sections we discuss these circuits in some detail to
see what their characteristics are and how their transfer functions are derived.

4-16.1 The Modulator

There are control systems designed to operate on either DC or AC signals. There


are also control systems that operate on AC signals in certain portions of their
systems and DC signals in other portions. Consequently, these systems must
have the capability of converting one form of signal to the other. This is where the
modulator comes in. A modulator circuit is used for converting DC type signals
to AC signals. Circuits that do this conversion are sometimes called inverters.
Unfortunately, the name modulator is somewhat misleading if you happen to
have some radio theory background. You know then that a modulator circuit is
Section 4-16 Auxiliary Servo Circuits 129

used for combining a lower frequency signal, usually an audio-rate signal, with a
much higher frequency carrier signal to prepare the lower frequency signal for
radio transmission. This is somewhat similar to what the modulator does in a
control circuit. In this case, the modulator converts a DC signal whose amplitude
varies quite slowly (due to varying command levels and responses) as compared
to a true formal AC signal, to an AC signal whose amplitude or frequency varies in
accordance to the amplitude of the DC signal.
There are a number of circuit designs that can be used for the DC-to-AC
conversion process, all of which work very well. Years ago, these circuits were
electromechanical in their operation and used what were called vibrator reed
circuits (see Figure 4-39). They were popular in the automotive industry for con¬
verting the then 6 VDC car battery voltage to a much higher AC voltage signal.
This higher voltage was then rectified, that is, converted to DC, to supply the high
DC voltage needs of the car radio’s vacuum tubes.

O High-voltage AC O

— cismms&msu—
Step-up Transformer

Figure 4-39 Vibrating reed circuit for


“chopping” a DC voltage for conver¬
sion to an AC voltage.

The modern-day control circuit modulator operates on roughly the same


principle as the old-fashioned vibrator circuit. Figure 4-40 shows a solid-state
chopper that does the same thing as the mechanical vibrator. The DC voltage
signal is alternately chopped by the two transistors, Q\ and Q2, called switching
transistors. They replace the mechanical switching that was performed by the
vibrating reed in our old vibrator circuit. The transistors are far more efficient,
and there are no moving parts to wear out as was the case in the older electrome¬
chanical circuits. In the solid-state chopper, or inverter as it is more frequently
called, the two transistors are alternately turned off and on allowing the input DC
voltage to flow first in one direction through the center-tapped winding, N\, and
then the other. This produces a square-wave signal (voltage VT in Figure 4-40)
which, in turn, produces an AC voltage across the output winding, N2. The
amplitude of this transformed signal will be directly proportional to the amplitude
130 Chapter 4 Transducers and Control System Components

\/0ut

Figure 4-40 A transistor switch modulator or inverter.

of the input DC voltage supplied to the inverter. The alternate switching action of
the two transistors is produced and controlled by the alternate saturation of each
half of winding N\. This, in turn, causes the center-tapped feedback transformer
winding, 7V3, to alternately bias the bases of Q\ and Q2 allowing one or the other
transistor to conduct. R2 and C\ encourage the circuit to begin oscillation at
startup of the system.
The transfer function for the modulator is the following:

£out ACY
K mod (4-35)
Em DCV

4-16.2 The Demodulator

The demodulator performs the opposite function of a modulator. Again, the radio
enthusiast, when he or she hears the term demodulator, probably thinks of a
receiver circuit that removes the low frequency audio signals from the incoming
Section 4-16 Auxiliary Servo Circuits 131

high frequency carrier radio signal. In the control circuit demodulator, somewhat
of the same process takes place. However, in this case, the amplitude or fre¬
quency of the AC signal containing the command level information is transformed
into a DC voltage whose amplitude is a representation of this command informa¬
tion.
If the demodulator converts the amplitude of the AC signal to a DC signal,
that circuit is called an amplitude demodulator or simply AM demodulator. If the
demodulator converts the frequency of the AC signal to DC signal, that circuit is
called a discriminator. The AM demodulator can be a simple diode that rectifies
the AC signal into a DC level signal. The filtering out of any remaining AC
components is done by R-L or R-L-C networks, as shown in Figure 4-41(a).
Figure 4-41(b) shows a full-wave demodulator that is more efficient than the

\
\
/ \
ACV DCV
\ I
\ /
\
y
/
\
y

D1 = Diode
/?1 C: C2 = PI Filtering Network for Reducing AC Components Occurring at Output
Rl = Output Load
(a) Half-wave Demodulator

\
\
DCV
I
/
/

D^D2D3Da = Diodes Arranged in Full-wave Bridge Configuration


/?., C1 C2 = PI Filtering Network
Rl = Output Load

(b) Fuil-wave Bridge-type Demodulator

Figure 4-41 Two popular demodulator circuits.


A. Half-wave demodulator
B. Full-wave bridge-type demodulator
132 Chapter 4 Transducers and Control System Components

circuit in Figure 4-41(a), producing a larger output signal as a result. FM discrimi¬


nator circuits dealing with servo applications are rather complicated in their cir¬
cuit design and are not discussed here.
The two circuits in Figures 4-41(a) and 4-41(b) represent the very simplest of
circuits that can be used. There are many other more sophisticated circuits that
won’t be covered here and would be superior in performance to the ones shown.
Nevertheless, the circuits presented here represent the very basic circuit princi¬
ples used in these applications.
The transfer function for a demodulator (regardless of whether it is designed
for converting AM or FM signals to DC) may be stated as:

Eout DCV
-•'-demod ^7. ACV

4-16.3 The Phase-Lead Network

The phase-lead network, as the name implies, is used to cause an existing phase
angle to become leading. This type of circuit is often referred to as a compensat¬
ing circuit. In other words, the circuit compensates for a lagging phase angle by
reducing the phase-lag which may be present between the output signal and the
input signal of a control system by shifting the output signal into a more leading
phase angle. As we find out in a later chapter, system stability is determined by
the phase relationship existing between the output and input signals. By shifting
this phase, we can stabilize or destabilize our control circuits at will. Figure 4-42
shows a typical phase-lead network. As you can see, the circuit is comprised
solely of passive components. No external power is required to operate the cir¬
cuit. The transfer function is derived by the following process:

Ri
-WW-

Out

o o

E 0ut = R2^ry + fi
E |n (/?, + R2)(ST2 + D
where r1 = /?•, C
/? i R2
r2 = —(C)
R) + R2 Figure 4-42 Phase-lead network.
Section 4-16 Auxiliary Servo Circuits 133

Looking at Figure 4-42 and using the voltage divider method for impedances:

Z,out R R
'in RjsC Rx + MsCRx + 1)
+ R:
1 sC sC
R\ +
sC sCR\ + 1
sC

R R2(sCR\ + 1)
Rx + sR2CR , + R2 Rx + R2 + R\R2sC
sCR i + 1
R2(sCR , + 1) R2 st i + 1
(4-37)
R\R2 R\ F Rl ST2 + 1
(R i + Ri) sC + 1
R\ + R
where tj = R\C
R\R2
T2
(Ri + Ri)C
Because of the resistive component, R2/(R\ + R2) in Eq. (4-37), there is an inser¬
tion loss as a result of using this circuit. Usually, this can be nullified by increas¬
ing the gain of an existing amplifier in the control circuit.

4-16.4 The Phase-Lag Network

The phase-lag network is used to do just the opposite from the phase-lead net¬
work. The purpose of the phase-lag network is to cause the existing phase angle
of a system to increasingly lag, thus compensating for a too leading phase angle.
Again, this network is comprised of passive components and therefore requires no
external power supply. Its circuit is shown in Figure 4-43. The derivation of its
transfer function is the following:

o- O

£,n E Out

o
EOux = 1 + *r1

£,n 1 + st2
where r1 = R2C
t2-(FI^+ R2)C Figure 4-43 Phase-lag network.
134 Chapter 4 Transducers and Control System Components

Again, using the voltage divider rule for impedances:

1 + SCR2
sC 1c
sCR \ + 1 + sCR2
R'+lc + Rl ~sC

1 + sCR2 1 + ST 1
(4-38)
1 + sC(R 1 + R2) 1 + st 2

where t\ = R2C
t2 = (R\ + R2)C

Servomotors and phase-lag networks have certain similar characteristics.


When the frequency response curves are plotted for both devices, there are close
similarities in the appearances of their response curves. As a result, servomotors
are sometimes referred to as phase-delay devices.

4-17 THE STEPPER MOTOR

The stepper motor is a type of motor that lends itself very nicely to the digital
controlling of automatic control systems. The stepper motor has found use in
many low torque applications where computer control is used. Good examples
are found in the design of printers and plotters.
Figure 4-44 represents a very basic depiction of what goes on inside a step¬
per motor. Shown is a four-pole motor having alternate pole polarities. The po¬
larity of each pole is determined by a set of digital bits, each bit determining the
polarity of its particular pole. In the example shown in Figure 4-44, a 1 bit causes
a north pole to appear at the pole’s end adjacent to the armature, while a 0 bit
causes a south pole to appear. By systematically switching the bit patterns in the
sequence shown in Table 4-2, it’s possible to create rotation in the stepper motor’s
armature, as seen in Figure 4-45.
Having established the method of rotation, it then becomes a matter of
developing the proper sequence of bits to cause the stepper to step in the desired
directions. This would become the function of the software program.
Stepper motors are manufactured to index in various amounts measured in
degrees. A stepper having a stepping angle of 45° would obviously have only 8
discrete positions in a full 360° angle of rotation. On the other hand, a stepper that
can step as little as 5° per index has 72 discrete positions, which means finer
resolution for positioning purposes.
The transfer function for a stepper motor, for the purposes of this text at
least, is similar to those for the servomotor. The same dynamic and transient
Section 4-17 The Stepper Motor 135

Q Pole 2 Q

O Pole 4 6

Figure 4-44 The “stepping” of a stepper motor.

TABLE 4-2 A STEPPER MOTOR’S


CONTROL BIT PATTERN FOR
ROTATION

Counterclockwise rotation
Pole 1 Pole 2 Pole 3 Pole 4

1 0 0 1
1 1 0 0
0 1 1 0
0 0 1 1

Clockwise rotation

0 1 1 0
1 1 0 0
1 0 0 1
0 0 1 1
a
’§C
■3
x
(U

_c
Ofi
’C4>
<—>
oj

03
E4>
•i—>
C/3
>4
C/3

oJ
>>
X)
4>
c
o
"O
c/3

4)
S-
3
■4—»
c3
E
3 oo
i—
-4—1

X4) E
■4—1
<+» u
o 4)
Cl
c CL
o 4>
■4—1 ■4—4
C/3
03
*-> 4)
o
s- X
4)
<U O
a. v
E g
O Cl
U xO
in a>o3
"T
rr +jj
0> on
tm 4)
3 3
M -3
•— 03
ta >

136
Summary 137

problems of motion are involved with the stepper as there are with the servomo¬
tor. The transfer function for a stepper is:

KV
KstM — rad/pulse (4-39)
5(1 + Ts)

For additional information on the stepper motor and how it operates, refer to
Section 12-4 in the last chapter.

SUMMARY

Transducers are essential components that make up a control system. Because of


their importance, we investigated their characteristics and determined the transfer
functions for many of the more frequently used sensors. We also discussed the
characteristics of the many servomechanisms commonly used in control sys¬
tems. We found that the servomechanism is a precision electromechanical device
used in feedback systems for creating the output commands of position and veloc¬
ity. We also investigated the synchromechanism and discussed its transfer func¬
tions. The synchromechanism is also a precision electromechanical device but is
used primarily in positional and positional measurement applications. It can
sometimes be used without amplifiers and is designed for analog applications.
Table 4-3 summarizes the transfer functions that were discussed in Chap¬
ter 4.

TABLE 4-3 SUMMARY OF TRANSFER FUNCTIONS

Device Transfer function Typical units

1. Potentiometer displace¬ F0ut volts/inch, volts/deg,


or ^f = Kp
ment transducer / or volts/rad
Fout
2. Velocity xducr volts/in/sec
vel = Kr
Fout FphotO
3. Photocell vel. transducer volts/ft/sec
vel 1 + ST

Fout Kot
4. Pressure xducr volts/lbs/in2
Pi n 1 + ST

J7
^out Fsonic
5. Microphone or sonic xducr volts/lb/in2
PdB 1 + ST
Fout Fflowrate volts/lb/sec or
6. Flowrate xducr
> Q 1 + 5T volts/ft3/sec
output volts
7. RF receiver Krf- volts//uvolt
input volts
J7
-^out Fthrm
8. Thermo xducr volts/°F
°F 1 + ST
138 Chapter 4 Transducers and Control System Components

TABLE 4-3 Cont.

Device Transfer function Typical units

Fout _ Xtcpi
9. Thermocouple mV/°F
°F 1 + ST

Fout — Fphoto
10. Photocell transducer volts/lumen
Q 1 + ST

11. Servomotor (VDSM) V


kvdsm - i(1
Kv
+ rad/volt

Kv( 1 + r2)
12. Servomotor (IDSM) K/D5M J(j + + STj) rad/volt

13. General TF for stepper K Ky


rad/pulse
motor Ks,M ~ s( 1 + TS)

^output _ p
14. Gear train n ~ Agear deg/deg
t'input

FoUt _ j'
15. Rate generator volt/rad/sec
Win

Fout is
16. CX or TX synchro n ~ A CX/TX volt/rad
“in

^OUt is
17. CR or TR synchro p ~ tSCR/TR rad/volt
Ain

R2 ST i + 1
18. Phase-lead network
R\ + R2 st2 + 1

1 + ST i
19. Phase-lag network
1 + ST 2

EXERCISES

4-1. Determine the transfer function for a potentiometer transducer system that
produces a total of 12.8 volts DC output for a total displacement of 42 inches
of its sensing arm.
4-2. Find the transfer function of a gear transmission system for an antenna
tracking system having the following specifications for the individual gears
(refer to Figure 4-32):

input gear 1 = 57 teeth

gear 2 = 108 teeth

gear 3 = 63 teeth

output gear 4 = 211 teeth


4-3. During the testing of a servomotor to determine its time constant, it was
found that it required 0.090 seconds for the motor to respond fully to a
References 139

command signal given it. Measurements were made with a recording oscil¬
loscope. What would be its actual time constant?
4-4. Referring to Figure 4-42, develop the transfer function for a phase-lead
network having the following component values:

R\ = 22 ohms

Ri = 1,000 ohms

C = 1,000 pF
4-5. In order to create a particular frequency response curve for a servo system,
a phase-lag network (see Figure 4-43) must be installed in which r2 must be
four times the amount of rj. If C is 1.3 /jlF and Rj is 100 ohms, what is the
phase-lag network’s complete transfer function?
4-6. Find the velocity constant for a servomotor whose no-load speed is 6,500
rpm and whose rated control voltage is listed as 32 VAC.
4-7. A certain manufacturer of VDSMs lists the following information for one of
its servomotor models:

time constant = 0.016 sec

no-load speed = 5500 rpm

rated control voltage = 24 VDC

Derive the servomotor’s transfer function based on the foregoing informa¬


tion.
4-8. A certain VDSM has a no-load speed of 377 rad/sec. Using Figure 4-30 as
its speed-torque curve, find its actual viscous damping amount in dyne-cm/
rad/sec.
4-9. A servomotor manufacturer lists the following transfer function for one of
its models:

25(1 + 0.03/co)
KIDSM
jo)( 1 + 0.05/co)(l + 0.004/co)

Correct this transfer function to show actual values.


4-10. In the following transfer function, calculate the actual phase angle and
magnitude of the output in degrees and rad/volt at a frequency of 12 rad/sec.

67(1 + 0.009s)
KIDSM
s( 1 + 0.007s)(l + 0.067s)

REFERENCES

Allocca, John A. and Allen Stuart, Transducers, Theory & Applications,


Reston, Va.: Reston Publishing Co., 1984.
140 Chapter 4 Transducers and Control System Components

Honeycutt, Richard A., Electromechanical Devices, Englewood Cliffs, N.J.:


Prentice-Hall, Inc., 1986.
Miller, Richard W., Servomechanisms, Devices & Fundamentals, Reston,
Va.: Reston Publishing Co., 1977.
O’Higgins, Patrick, Basic Instrumentation, New York, N.Y.: McGraw-Hill
Book Co., 1966.
Patrick, Dale R. and Stephan W. Fardo, Industrial Process Control Sys¬
tems, Englewood Cliffs, N.J.: Prentice-Hall, Inc., 1979.
5
Block Diagram
Mathematics

5-1 BREAKING UP SYSTEMS INTO BLOCKS

Admittedly, discussions of block diagrams or flow charts are not the most dy¬
namic portions of any textbook. These figures are typically used for explaining
the operation of some kind of system. Or they may be used to show the flow of
information within a system, showing how it gets from point A to point B. In
automatic control systems, block diagrams are used for many of the same kinds of
applications. However, there is an added dimension to their usage. There exists
a rigorous, math-disciplined method for changing and moving the locations of
these blocks within a system; this is usually done for the purpose of making
system simplifications.
In an automatic control system, each block represents a system component
having a particular transfer function. And, it’s these transfer functions that must
work with the other transfer functions in an orderly systematic fashion so as to
create the desired overall system output. Look at Figure 5-1. Here we see a
block diagram of a somewhat simplified version of an automobile’s power plant
system. Study it for a moment. Not the inputs. These are the driver’s inputs
needed to properly control the car’s various systems. Remember too, that like the
simple block diagram systems that we studied back in Chapter 2, each of the
system components seen in Figure 5-1 has its own unique transfer function.
For example, look at the block labeled Transmission in Figure 5-1. The
transfer function for this system component would be output/input = OouJ 0in ,
where 60ut represents the angular displacement of the output shaft and 0in is the
angular displacement of the input drive coming from the engine. If we were to

141
142 Chapter 5 Block Diagram Mathematics

Figure 5-1 A block diagram of the power plant in an automobile.

swap the transmission for, say, the brakes system, our intuition would tell us that
we would have a disaster on our hands if we attempted to drive that car. But let’s
look at this problem mathematically. The transfer function for the braking system
would look something like this: pressure/displacement. Pressure is the output
hydraulic pressure needed to operate the wheels’ brakes; displacement is the
brake pedal movement needed to create the pressure. The transfer function is
obviously radically different from that of the transmission’s. If we investigated
the other system components in our car, we would find a similar lack of input-
output compatibility. In other words, if we were to swap components around in
our car, we would find that each component has been specifically designed to
operate in that, and only that, location or sequence of systems. And if we were to
substitute or combine components, we would want to make certain that the inputs
and outputs of the changed components match exactly with the systems they must
work with. This is because we want our car to have the same output performance
that we’ve been accustomed to experiencing.
An automotive design engineer, when designing a new car system, has a
very good idea what transfer functions are required for each component, in the¬
ory, to make the car perform properly. If he or she decides to combine or swap
these components for the sake of simplifying the system, the output versus input
results must remain the same. These must not change.
Section 5-1 Breaking Up Systems into Blocks 143

Let’s look at another system now. Figure 5-2 shows a drawing of a system
that can be used for automatically tracking the trajectory of a satellite-launching
rocket, for photographic purposes. A series of infrared photocell sensors are used
for following the heat exhaust of the rocket. The sensors cause servomotors to
continuously adjust the telescope and camera assembly for proper attitude. Fig¬
ure 5-3 is the block diagram of this system. Each system component block has its
transfer function written inside.
As in the case of the automotive engineer, the design engineer on this partic¬
ular project would want the freedom to replace blocks, combine blocks, add and
remove blocks, while still maintaining a desired overall transfer function for that
portion of the system being worked on. Bear in mind too, that merely changing a
single wiring or piping connection to one of these blocks can greatly affect the
system’s overall transfer function. The point is this: In order to have this freedom
of design change, a thorough knowledge of block diagram mathematics is neces¬
sary.

/
y
y

Figure 5-2 An automatic telescopic camera for tracking rocket trajectories.


144 Chapter 5 Block Diagram Mathematics

Transfer Function
of Photocell and
Optics = 1.23 V/Lumen

Transmission Mechanical
Turret Tracker

Figure 5-3 Block diagram for the automatic telescopic camera described in Fig¬
ure 5-2.

5-2 REVIEWING OLD TERMS AND LEARNING NEW ONES

In order to understand how block diagram mathematics work, we must first under¬
stand the symbols used in block diagrams. Refer to Figure 5-4. The four symbols
shown are defined as follows:

1. The block: The block represents a component, or a combination of several


components, of a system. The block has a signal input and an output.
Associated with this block is its transfer function. Usually, all power con¬
nections to this block are omitted simply for clarity reasons.

The Block

The Directional Arrow-►

The Take-off Point

The Summing Point

A I d, (As Shown), Positive Feedback Being Used


I If ", Negative Feedback Being Used

Figure 5-4. The four major symbols used in automatic controls block diagram¬
ing.
Section 5-2 Reviewing Old Terms and Learning New Ones 145

2. The directional arrow: The directional arrow indicates the direction of flow
of a signal or data into and out of each block or along a circuit line.
3. The take-off point: This is a point where two or more circuit lines are physi¬
cally joined together allowing their separate signals to combine into one
signal, or to allow a single signal to split into the separate lines. In the
second case, it is assumed that the divided signal is of the same amplitude
and phase as the original signal.
4. The summing point: This is a point where two signals are either added
together or subtracted from each other. Note that despite the fact that two
signals may be subtracted from each other, it is still called a summing point.
Note that if bottom sign is (+), positive feedback is being used in the system;
if bottom sign is (—), negative feedback is being used.

Let’s review a few of the basic points mentioned in Chapter 2 concerning


system gain, positive feedback, and negative feedback:

.
1 System gain: May be considered another way of saying transfer function;
used primarily in electronics to describe the amount of amplification of an
amplifier or amplifier system. It is the ratio of the voltage or current output
divided by the system’s voltage or current input.
.
2 Positive feedback: A condition found in certain control systems where the
output is coupled back to the input, in phase, so that there is continuous
reamplification of the original input signal. Mathematically, this is described
as:

'Out
(see Eq. 2-8)
Em 1 - AB
3. Negative feedback: A condition used frequently in control systems where
the output is coupled back, out of phase, to the input in order to cancel a
portion of the input signal. Mathematically, this is described as:

'Out A
(see Eq. 2-14)
'in 1 + AB

The following is a list of eight basic rules or identities for changing block
diagrams for the purpose of simplifying or modifying them. These rules are stated
in the form of block diagrams and should be studied closely for their understand¬
ing.

1. Cascading blocks: Refer to Figure 5-5.

n. n
u2 c w (Jl Cj2
ui * °Out °ln

Figure 5-5 Cascading blocks.


146 Chapter 5 Block Diagram Mathematics

2. Eliminating a forward loop: Refer to Figure 5-6.

rz + n
'Out ^ln °1 1 ^2 Out

Figure 5-6 Eliminating a forward loop.

3. Moving a take-off point from an input to an output of a block: Refer to Figure


5-7.

Figure 5-7 Moving a take-off point from an input to an output of a block.

4. Moving a take-off point to a block ahead: Refer to Figure 5-8.

^Out ^In S Out

'Out S Out

Figure 5-8 Moving a take-off point to a block ahead.

5. Moving a summing point beyond a block: Refer to Figure 5-9.

Figure 5-9 Moving a summing point beyond a block.


Section 5-3 Cascading Blocks 147

6. Moving a summing point from an output to an input of a block: Refer to


Figure 5-10.

Figure 5-10 Moving a summing point from an output to an input of a block.

7. Eliminating a feedback loop: Refer to Figure 5-11.

G
Out = 5, - W
Out
1 + HG

Figure 5-11 Eliminating a feedback loop.

8. Switching blocks: Refer to Figure 5-12.

s Out

Figure 5-12 Switching blocks.

With these rules, you can combine blocks into single blocks, move pick-off
points, or shift summing points ahead or behind blocks. Let’s try our hand now at
using these rules by going through some examples.

5-3 CASCADING BLOCKS

Probably the most often encountered system configuration in any automatic con¬
trols circuit is the one in which the output of one system is fed directly into the
input of the next system, and so on. This is called a cascading system. Rule 1
148 Chapter 5 Block Diagram Mathematics

shows how this type of configuration is simplified or reduced to a single block. As


shown by this rule, the transfer function of the simplified version is nothing more
than the product of all the individual blocks’ transfer functions. The following
example illustrates this point.

EXAMPLE 5-1
Simplify the system in Figure 5-13 by using rule 1.

1 In ’ Out

Figure 5-13 Example 5-1. The figures in each block are the individual transfer
functions and Sm and 50Ut are the input and output signals for the total system.

Solution:
According to rule 1, the equivalent simplified system would have a transfer function
equivalent to

12.4 x 6.3 x 0.87 x 23.7 = 1610.76

5-4 ELIMINATING LOOPS

Often, it’s desirable to get rid of feedback loops for the sake of simplification and
cost reduction. This can be done as long as the feedback loop doesn’t need any
adjustments. If it can remain fixed in value, then most likely it should be simpli¬
fied. The following example shows how this can be done.

EXAMPLE 5-2 -
Simplify the system in Figure 5-14 by finding the equivalent transfer function.

Figure 5-14 Example 5-2.

Solution:
The system shown in Figure 5-14 matches the configuration in rule 7. Note that
Figure 5-14 is a negative feedback circuit. Therefore,
Section 5-5 Moving Summing Points 149

‘Jput _ 7

Sm - \ + GH
Since G = 115 and H = 11, then

Sout = 115
5in 1 + (11X115)
= 0.091
Let’s see what happens now if we change the circuit in Figure 5-14 to a positive
feedback circuit as shown in Figure 5-15 and try to simplify it.

EXAMPLE 5-3 -
Simplify Figure 5-15.

’ Out

Figure 5-15 Example 5-3.

Solution:
Notice that the polarity sign at the bottom of the summing junction is now reversed to
(+). Again, using rule 7, but being careful now to use the proper sign in the transfer
equation:

'out G
S;in 1 - GH
Since G = 115 and H = 11, then
S out 115
.Sin 1- (11X115)
- -0.091
Since there is no such thing as a negative gain or transfer function, the preced¬
ing results must be discarded. The system, in other words, is unworkable and has to
be redesigned.

5-5 MOVING SUMMING POINTS

At times it is advantageous to move summing points from one location to another


in a circuit because of physical space requirements or because of other design
limitations. The following two examples show how this can be done.
150 Chapter 5 Block Diagram Mathematics

EXAMPLE 5-4
In Figure 5-16, the summing point is to be moved to the output side of the block.
Find the new equivalent system.

S Out

Figure 5-16 Example 5-4.

Solution:
This type of configuration is described by rule 5. Figure 5-17 shows the solution.
Notice the addition of an identical block to the original circuit block in the Sin2 line
going to the newly moved summing point.

Figure 5-17 Solution to Example 5-4.

EXAMPLE 5-5 ---


Refer to Figure 5-18. Assume that we want to move the given summing point to the
input side of the given block. What is the resulting new equivalent circuit?

^^ +

S,n2 = 2.02 Figure 5-18 Example 5-5.

Solution:
The solution to this problem is seen in Figure 5-19. Notice the addition of a block in
the newly relocated Sin2 line whose transfer function is the reciprocal of the originally
given block. The magnitude of 5oUt = (0.85)(32) + 2.02 = 29.22.

1 Out

Figure 5-19 Solution to


Example 5-5.
Section 5-7 Combined Block Reductions 151

5-6 INTERCHANGING BLOCKS

The following example shows how you can swap the positions of two blocks with
each other without changing the overall system transfer function.

EXAMPLE 5-6 -
Referring to Figure 5-20, we would like to interchange the two blocks without chang¬
ing 5out/5jn.

■^Out _ 15
Sln 1 + (3) (15)
= 1!L
46

= 0.326 Figure 5-20 Example 5-6.

Solution:
The solution is shown in Figure 5-21. Notice that the system transfer function for the
interchanged system did not change.

Sln 1 + (-)(—)
3 ' 15 '

0.333
1 + 0.0222
= 0.326 Figure 5-21 Solution to Example 5-6.

5-7 COMBINED BLOCK REDUCTIONS

We now look at some complex block systems requiring several combined changes
and reductions in order to eventually wind up with a much simplified system.
There are many instances in which we won’t want to reduce the number of blocks
152 Chapter 5 Block Diagram Mathematics

in a given system. There are times when it’s more desirable to have individual
blocks of systems rather than having them combined into one or just a few. From
the standpoint of servicing a complex system, many times it’s far cheaper to be
able to remove an individual block for replacement rather than replacing a more
complex circuit or block that contains many subcircuits or assemblies. The de¬
sign decision must be made with some marketing knowledge for that product
design.
Let’s try our hand at reducing the system shown in Figure 5-22. The idea
here is to reduce the system down to just one block and its transfer function.
Again, bear in mind that from a practical standpoint, this may not be the most
desirable thing to do. This is simply an exercise in block diagram mathematics.

EXAMPLE 5-7 -
Reduce the system in Figure 5-22(a) to a single block and its transfer function using
the block reduction math method.

(a)

Figure 5-22 Example 5-7.

Solution:
The solution is shown in Figure 5-22(b) through (g) and summarized in the following
step-by-step instructions. These steps can be applied to a majority of all block dia¬
gram math problems:

1. Move any blocks that are outside of any loops to the inside of these loops;
Figure 5-20(b).
2. Combine any cascaded blocks; Figure 5-20(c).
3. Combine loops within loops where possible, when they share the same sum¬
ming points and take-off points; Figure 5-20(d).
4. Combine any cascaded terms again where necessary; Figure 5-20(e).
5. Combine any remaining loops into one loop; Figure 5-20(f).
6. Finally, eliminate the remaining loop; Figure 5-20(g).
Section 5-7 Combined Block Reductions 153

Out

(b)

(c)

Sin

Combine Summing
Points (Rule 6)

(d)

Figure 5-22 (cont.)


154 Chapter 5 Block Diagram Mathematics

Combine Cascaded
Terms Once Again
Where Necessary Out
(Rule 1)

(e)

Combine Any Remaining Loops


into One Loop. (This Can Be
Done by Noting That Loops 1
and 2 in (e) Merely Subtract from
Each Other. Subtract Loop 2
from Loop 1 to Keep Expression
Positive Rather than Negative.)
1 _ H = Gi ~ H
G3G4 G1G3G4 GiG3G4

Eliminate Remaining Loop by


Using Rule 7. Since This is a
Positive Feedback Loop, Use Gi G2G3G4
G 1 In Out
5)

3
I

1 - HG
Gy~ H
Note: The G Term in the Positive Feedback Loop is Gy G2G3G4 and the H Term is -
Gy G3G4
Therefore:

Gi G2G3G4 Gy G2G3G4
Gy~ H 1 - (Gy - H)G2
(G, G2G3G4)
Gy G3G4
(g)

Figure 5-22 (cont.)

EXAMPLE 5-8

Determine the magnitude of the output signal, Sout> in Figure 5-23(a), using the block
reduction math method.
Section 5-7 Combined Block Reductions 155

(a)

Figure 5-23 Example 5-8.

Solution:
The solution is presented in Figure 5-23(b) through (e).

(b)

’ Out

(c)
156 Chapter 5 Block Diagram Mathematics

Out

S,n = 2.0 ’ Out


= ?

■^Out _ ^Out _ 60
S,n 2.0 1 + 0.84) (60)
= 1.167
S0ut = (1.167M2.0)
= 2.334

(e)

After working a number of block reduction problems, you will undoubtedly


find numerous shortcuts that can be taken in solving these kinds of problems.
Rather than presenting these shortcuts here, it’s better that you find them on your
own. That way, you’re less likely to forget them.

SUMMARY

Block diagram mathematics can be a very worthy time-saving exercise to go


through for system simplification. The amount or degree of simplification de¬
pends on what system features are needed. The rules for block diagram math
were summarized in Figures 5-5 through 5-12.

EXERCISES

The following exercises are strictly exercises in theory, applying the information
discussed in this chapter. These are not problems that represent actual function¬
ing systems.
5-1. Reduce the block diagram in Figure 5-24 to a single transfer function using
the techniques outlined in this chapter.

20 50 35 E Out
Figure 5-24 Problem 5-1.
Summary 157

5-2. Reduce the block diagram in Figure 5-25 to a single transfer function, again,
using the techniques that were discussed.

E n

Figure 5-25 Problem 5-2.

5-3. Do the same for Figure 5-26.


+ 10
\
I
W
> E Out
s + 5

Figure 5-26 Problem 5-3.

5-4. Do the same for Figure 5-27.

i 8
E n
_^
^ \X Out
s + 1 s + 5
i i —

20

Figure 5-27 Problem 5-4.

5-5. Do the same for Figure 5-28.

Out

Figure 5-28 Problem 5-5.


158 Chapter 5 Block Diagram Mathematics

5-6. Do the same for Figure 5-29.

Figure 5-29 Problem 5-6.

5-7. Do the same for Figure 5-30.

Figure 5-30 Problem 5-7.

5-8. Do the same for Figure 5-31.

Figure 5-31 Problem 5-8.


Summary 159

5-9. Do the same for Figure 5-32.

Figure 5-32 Problem 5-9.

5-10. Do the same for Figure 5-33.

In • Out

Figure 5-33 Problem 5-10.


6-1 INTRODUCTION

Hopefully, this chapter will open up a whole new perspective for you for under¬
standing automatic control systems. Up to this point, most of what has been said
on this subject has been ground work material. That is, with the possible excep¬
tion of Chapter 2, virtually all that’s been discussed up to now has had to do with
preparing you for understanding automatic control concepts. To help us venture
into this new territory, we first study the concept of the Bode plot. The Bode plot
is to the automatic controls engineer what the frequency response curve is to the
stereo enthusiast. Both plots tell the person what kind of gain can be expected out
of his or her system for a particular given frequency input.

6-2 BODE DIAGRAM CONSTRUCTION

H. W. Bode (pronounced, Bo-dee) worked for the Bell Laboratories following


World War II. Bode developed a systematic method of mathematically analyzing
a controls system and plotting the results on graph paper. The method of presen¬
tation on the graph paper is what makes the Bode method so unique and easy to
understand. In order to appreciate the Bode plot, let’s first review the discussion
in Chapter 1 concerning frequency response.
We described a method in Section 2-4 on how to observe the linearity of an
audio amplifier. In place of the amplifier, however, we could have used virtually
any kind of audio circuit to see how the output varied with its input as the input

160
Section 6-2 Bode Diagram Construction 161

frequency is varied. As a matter of fact, if we knew the transfer function of the


audio circuit being tested, we could have plotted the results ahead of time. We
wouldn’t have had to have gone through the trouble of setting up the hardware at
all! And this is exactly what we want to be able to do with the circuits we study
here. We want to be able to plot the transfer function characteristics based solely
on the transfer function itself. We don’t want to have to set up a circuit and
record measurements. That’s too time-consuming.
Let’s take another look at our test circuit in Figure 2-3. In place of the audio
amplifier, we substitute a resistor network, shown in Figure 6-1. Let’s assume
that this circuit has an input, a-b, and an output, c-d. Our dual-trace oscillo¬
scope is also hooked up across these connections as shown. What we want to do
now at this point is to calculate the transfer function of this circuit.

Figure 6-1 A test setup for determining the transfer function of a resistor net¬
work.

Look now at Figure 6-2. This is the very same circuit as seen in Figure 6-1,
but the resistors in the network have been shifted around slightly so that the input
and output voltage relationships can be seen a little more clearly. We can also see
that resistors Rx and R2 will have no affect on the transfer function relationship
since they don’t determine the value of em. In other words, em has already been
fixed in value, regardless of the input resistance. Of course, we are assuming that

Figure 6-2 The circuit in Figure 6-1,


redrawn in schematic form.
162 Chapter 6 Bode Diagrams

ein is being supplied from a power supply capable of supplying any voltage at any
current demand placed on it by the input resistance of the test circuit. Now, using
the voltage divider rule (see Section 1-5), we see that

^out
= K network
Gn Rt, + R4

which now becomes our transfer function for this circuit.


The next thing we want to do is to plot the frequency response curve for our
circuit. This will help us to understand the mechanics of plotting when we attempt
to plot the more complex circuits associated with a controls system. And this
brings us to our first rule in plotting frequency responses in a Bode plot: All
amplitudes are expressed in decibels rather than in voul/vm terms. In other words,
rather than plotting frequency versus a voltage ratio, we will, instead, plot fre¬
quency versus decibels. The reason for this is: In many cases components are
cascaded, such as in the case of two or more amplifiers connected in series with
each other. The overall frequency response curve of this system will be equal to
the product of the individual frequency response curves, assuming that we were
plotting our results on a voltage ratio versus frequency graph. However, multi¬
plying curves together is not the easiest thing to do graphically. But if we were to
use the decibel system instead, all that is needed is to add the individual curves
together. This will then give us the final overall curve for the system. Then, if we
wanted, we could convert the final curve results back into the voltage ratio do¬
main by taking the antilog of our results.
This brings us to our second rule in Bode plotting, which concerns fre¬
quency. Many times, but not always, the frequency is expressed in radians/sec
rather than in hertz. The conversion from the one system to the other is simple:
a) = 2irf. Unfortunately, the gj is still referred to as frequency by many controls
engineers, which may make a few non-controls engineers cringe. Obviously, the
(o is an angular velocity term. However, custom takes precedence here. The
reason for the frequency-to-rad/sec conversion is due to the frequent occurrence
of the jco term in controls mathematics. In this book we use co for our frequencies.
Looking at the transfer function that we obtained for our foregoing resistive
network, you have probably guessed by now that we are dealing with nothing
more than a pure number, a constant. And when we take the log of this number
we will still come up with a pure number. Therefore, we have a transfer function
that is frequency independent. That is, regardless of frequency, we will always
have the same value of amplitude. So we must keep this in mind when we eventu¬
ally plot this information. And speaking of plotting, since we’re dealing with logs
here, all of our plotting will be done on semilog paper. That is, the x axis, or
abscissa, will be comprised of logarithmic-spaced lines. On the other hand, the y
axis, or ordinate, will be formed by evenly spaced or rectilinear lines. The units
associated with this axis will be decibels. As a matter of fact, you will often see
the designation dBA being used. This means, “Gain (A) converted to decibels.”
Recalling the purpose of log paper, any exponential curves existing in our transfer
Section 6-2 Bode Diagram Construction 163

functions will generally stand a better chance plotting as a straight line, or nearly
so, than if they were plotted on rectilinear paper. The one thing we must keep in
mind though is the fact that log paper comes in various cycles to suit different x
axis data ranges. To select the most suitable paper, determine the number of
decades or powers of 10 that your x axis data covers. If, for instance, your data
ranges from a value of 0.7 to, say, 3,400, then you must use five-cycle paper. This
is because you will be plotting data from 0.1 to 1, 1 to 10, 10 to 100, 100 to 1,000,
and 1,000 to 10,000.
The third rule in developing Bode plots is the often-done practice of plotting
phase angle information on the same graph paper as the frequency versus ampli¬
tude information. (To refresh your memory on phase angles, refer back to Section
2-5.) Since our transfer function in Figure 6-2 is not frequency dependent in any
way, the phase shift between the output and the input of our resistor circuit will be
zero.
We are ready to make a Bode plot of our resistor network. But before we
can do this, we must assign some resistor values to the resistors in our circuit in
Figure 6-2. Let R3 = 40011 and R4 = 2000. Because Knetwork equals eonJem which
equals 200/(200 + 400), or 0.33, and this value will be the same regardless of
frequency (i.e., regardless of co), and since 20 log i0( ^network) equals -0.48 dB, our
frequency response, when plotted, will look like the curve in Figure 6-3. And
because the phase angle is the same (0°) regardless of frequency, the phase angle
curve of our Bode plot will also appear as shown in the upper half of our graph

180
ID
90 Ld
PHASE ANGLE,0=O’ 1=1
0 'w'

-90 u
_i
-180 u
z:
<r
Ld
c/)
<n
i
CL

c
pq
Z5
0
K=0,33( -0.48dB)
TO
<C
LD ■20

,1 10 100 1000
(Rad/sec)
Figure 6-3 The Bode plot of Figure 6-2 where R3 = 400 Cl and R4 = 200 ft.
164 Chapter 6 Bode Diagrams

paper. The results seen in Figure 6-3 are telling us that for any frequency indepen¬
dent component (i.e., there are no co terms in the transfer function), the Bode plot
of the gain has a slope of zero. Or, to state it in other terms, the slope on the Bode
plot is 0 dB per each decade as seen and plotted on logarithmic graph paper. Also,
the characteristic of the phase angle curve for this same kind of component is also
a straight line drawn at 0°. Figure 6-4 shows the Bode plots for several other
values of K.

ID
Ld
180 n
90 u
^FDR ALL VALUES DF K _J
0 UD
.
30 -90 <c
K=10 —-
20 -180 Ld
K=3,162 - (/)
GAIN (dBA)

10 <r
K=1.0 — JZ
0 Cl
K=0,3162 -
-10
K=0.1 —^
-20
K=0.013 —-
■30
■40
1 10 100 1000
(Rad/sec)
Figure 6-4 Bode plot for several values of the constant, K.

Now let’s explore a different component configuration. Again, we use elec¬


trical components for our experiments, but keep in mind that we could just as well
be using electromechanical components such as servomotors, generators, etc.
It’s just that electrical components such as resistors, capacitors, or inductors are
easier to set up mathematically and to envision what is happening. Later, we
investigate applications involving the electromechanical devices, but for now,
let’s take a look at Figure 6-5. Here we see another circuit but with a frequency
dependent component added, a capacitor. The transfer function for this circuit is:

[S ^out ^out

^rc ~ ecin Z
-^in

Again, we have used the voltage divider rule to come up with our solution,
but notice that in place of R we have now used impedance, Z. This, of course, is
Section 6-2 Bode Diagram Construction 165

R
o-vwv O

e In

O o Figure 6-5

necessary since we are now dealing with AC voltages with frequency sensitive
components (i.e., C). Using C to determine Z, our transfer function now becomes

_Zout
&rc ~ ~7~
^in

1
jcoC
1
R +
jcoC
When deriving transfer functions for Bode plotting, the numerator of that
function is usually forcibly set to 1 to make the plotting easier. We consider the
reason for this later. Therefore, we now multiply the numerator by jtoC, and of
course, in order not to change the value of the expression, we do the same thing to
the denominator:

jcoC
jcoC

R + jcoC
jcoC
1
(6-1)
1 + jcoRC

Or, more generally,

= -———, since r = RC (6-2)


1 + J(x)T

We refer to the form of Eq. (6-2) from now on as the standard form of a Bode
transfer function.
Now let’s assign some values to the variables to Eq. (6-1). Let R = 1 Meg
and C = 0.05 /zF. This will make RC = r — 0.05 sec. Equation (6-1) now be¬
comes:
1
(6-3)
1 + jcoO. 05
166 Chapter 6 Bode Diagrams

Our next chore is to plot Eq. (6-3) on our semilog paper. But to do this we
must first assign a wide range of values to w, and for each assigned value calculate
the gain, A, in dB. To do all of this and not lose track of what it is we are doing,
we construct a table to record all our values. Our calculations will look something
like this (refer to Appendix A to review converting rectangular complex number
forms, i.e., /-operators, to polar forms):

1 1 1
1 + j(OT Vl2 + (cot)2 COT
1 A A arctan
T
1 _ 1 A -6
A A 6 ~ A

Table 6-1 shows the results of our labor. We selected the values for oo that
would give us a fair distribution of points across our graph paper.
Now, let’s plot our figures, once again following the rules that were dis¬
cussed earlier regarding making Bode diagrams. Figure 6-6 shows the results.
Even though our table lists frequencies only in the range of 1 to 1,000 rad/sec for
ct>, we can assume that the curves resulting from our graphing extend to infinity in
both directions. In other words, there are no additional significant features that
occur beyond either curve end.

TABLE 6-1 RESULTS OF PLOTTING THE FUNCTION: -———


1 + jcj 0.05

CO 1 + 76)0.05 1/04 /I 6) dBA = 20log(l/A)

rad/sec result, exp. A A0 1/A ML6 dBA

1 1 + j'0.05 1.001 2.9° .999 A -2.9° -0.009


2 1 + J0.1 1.005 5.7° .995 A -5.7° -0.04
5 1 + j 0.25 1.031 14.0° .970 A —14.0° -0.26
8 1 + jO.4 1.077 21.8° .929 A-21.8° -0.64
10 1 + j 0.5 1.118 26.6° /. 894 A—26.6° -0.97
15 1 + 70.75 1.250 36.9° .800 A—36.9° -1.94
20 1 +j 1 1.414 45.0° .707 A-45.0° -3.01
30 1 +jl.5 1.803 56.3° .555 A-56.3° -5.11
40 1 + j2 2.236 63.4° .447 A—63.4° -6.99
50 1 +j 2.5 2.693 68.2° .371 A-68.2° -8.61
80 1 +j 4 4.123 76.0° .243 A—76.0° -12.3
100 1 +j 5 5.099 78.7° .196 A—78.7° -14.2
200 1 +j 10 10.050 84.3° .100 A-84.30 -20.0
400 1 + j 20 20.020 87.1° .050 A—87.1° -26.0
600 1 + 730 30.020 88.1° .033 A—88.1° -29.6
800 1 + 740 40.010 88.6° .025 A-88.60 -32.0
1,000 1 + j 50 50.010 88.9° .020 A-88.90 -34.0
Section 6-2 Bode Diagram Construction 167

Let’s take a closer look at Figure 6-6 now that we have constructed our first
Bode plot containing a frequency dependent component. Figure 6-6 is the Bode
plot for our RC network described in Figure 6-5. Looking at the phase angle curve
in Figure 6-6, we are actually looking at the phase shift relationship existing
between the output voltage, eout, and the input voltage, ein. It’s important at this

RAD./SEC.
Figure 6-6 Bode plot for the function 1/1 + jco0.05.

point to remember not to confuse this phase shift relationship with the current and
voltage phase shift relationship that exists in a purely capacitive reactance cir¬
cuit. We already know that regardless of capacitor size and frequency being used,
the current always leads the voltage by exactly 90° for a capacitor. However, in
our situation described in Figure 6-5 and Figure 6-6, we are dealing with impe¬
dance, the combined effects of resistors and a capacitor in a circuit. And further¬
more, we are comparing a circuit’s output voltage to its input voltage, not merely
a capacitor’s output voltage to it’s output current. It’s very easy to confuse these
phase relationships.
There are three significant observations that we make as we inspect the two
curves in our Bode plot. First, notice the rather sudden bend in the gain curve
that takes place around the 20 rad/sec segment of the curve (Figure 6-7). The
bending is in the range of 10 to 30 rad/sec, but the center occurs at approximately
20 rad/sec. This sudden bending is an event that occurs in many Bode gain plots,
and it can occur anywhere along the gain plot.
168 Chapter 6 Bode Diagrams

S~\

ID
n hi
- —r* i

- -HD
U 111
—'
CD
7 cr c
/D <£
+10 i nn Qj
1UU
in
m 0 d
X
~ -10 Q_
CENT ER c
1 -20 □F Q
^ -30 BEND o
OJ' - 1 DECADE
-40
- 1 DECADE - - 1 DECADE— AiNALYZING THE
‘ I3DDE PLEIT GIF
i ntL r u inu i i LI IN
1
l+iw.05
1 10 100 1000
RAD,/SEC,

Figure 6-7 Analyzing the Bode plot of the function 1/1 + jcoO.05.

The second observation has to do with the phase angle curve. Looking again
at the bending of the gain curve at the 20 rad/sec point, draw a straight line up to
where it intersects the phase angle curve. Notice that it meets the phase angle
curve at a value of -45°. From this point, look one decade to either side of this
intersection point. Notice that we encompass most of the S-shaped portion of the
phase angle curve. On either side of our one-decade zone we notice that the phase
angle curve is relatively flat and is close to 0° on the left side of the zone, and is
close to -90° on the zone’s right-hand side. This observation furnishes us with a
valuable tool for helping us to construct our Bode plots later.
The third observation again concerns the gain curve. Preceding the bend in
our curve, we notice the levelness of the curve. However, following the curve’s
sudden downward bend, we notice that the slope remains at a relatively constant
value. If we measure this slope, we find it to be about -20 dBA/decade (again,
refer to Figure 6-7).
Let’s go back now to our first observation concerning the bending point in
our gain curve. We had found that this occurred at 20 rad/sec. Look again at our
equation (Eq. 6-3), and notice the value of our time constant, r. By taking the
reciprocal of r, (1/0.05), we find that we come up with the value of 20. This is the
value of our gain curve’s bend center. When working with Bode plots, this bend,
and its value, is referred to as the gain curve’s corner frequency. This is the point
beyond which the frequency response of the system ceases to be one value and
gradually begins assuming another value.
Section 6-2 Bode Diagram Construction 169

Let’s now see if we can simplify our plotting chores in Figure 6-6 so that we
don’t have to construct a table like Table 6-1 each time we want to construct a
Bode plot. Instead of plotting individual points as we did in Figure 6-6, we can
now make these observations:

1. The corner frequency of our gain plot will occur at 20 rad/sec (1/0.05 = 20).
2. Our gain curve will be flat at a value of 0 dBA from this corner frequency
extending to its left.
3. At the corner frequency and to its right, the gain will slope downward at a
constant rate of -20 dBA/decade.
4. Looking at our phase angle curve, we can approximate the left-hand portion
of the curve with a straight flat line at 0° that extends from the far left end of
the plot to within one decade of the corner frequency found on the gain
curve.
5. Now, going to the extreme right-hand side of the phase angle plot, we can
draw another straight flat line at 90° that extends toward the left to, again,
within one decade of the gain curve’s corner frequency.
6. The central portion of the phase angle curve can be filled in with a straight
line that joins the two ends of the flat lines forming the two ends of the phase
angle curves. The straight-line approximations of both the phase angle
curve and the gain curve for our transfer function are shown in Figure 6-8.

ID
0 Ld
f=l
-25
-50 ^
CD
"75 £
-100 *
1/1
C5
JZ
CL

1 10 100 1000
RAD,/SEC,

Figure 6-8 The corner frequency.


170 Chapter 6 Bode Diagrams

Let’s now try another example to help us become familiar with this straight-
line construction process:

EXAMPLE 6-1
Construct a Bode plot, using straight-line approximation, for the function,

1
1 + jojO.015

Solution:
Step 1. We note that the transfer function is already in the form of Eq. (6-2).

Step 2. Determine the function’s corner frequency. This is done by finding the
reciprocal of the function’s time constant, r. Since r = 0.015 sec, the corner fre¬
quency is = 1/0.015 = 66.7.

Step 3. Lay out coordinates on three- or four-cycle semilog paper similar to the
ones used in Figure 6-6.

Step 4. Locate the corner frequency on the graph and extend a straight line from
this point to the extreme left of the graph. To the right of this point extend a straight
line sloping downward having a slope of -20 dBA/decade and extend it to the far
right of the graph. This completes the gain portion of the Bode plot.

Step 5. Draw a straight line at 0° from the extreme left side of the graph over toward
the center of the graph. Draw this line lightly. Draw a straight line at -90° from the
extreme right side of the graph over towards the center. Extend both lines so that
each comes within one decade of the corner frequency. It may help to extend a
lightly drawn vertical construction line upwards from the corner frequency on the
gain plot to help in locating the decade widths on the phase angle curve construction.

Step 6. After determining the proper lengths of the ends of the phase angle lines,
one being higher than the other, both horizontal, and having the center span missing,
draw in the central missing portion. This is done by merely connecting the two inside
line ends with a sloping straight line. The final results should look like Figure 6-9.

0 ID
UJ
n
-25
-50
0)
-75
-100

Figure 6-9 Straight-line approximation


curve for 1/1 -I- j'co0.05.
Section 6-3 The Elements of Bode Diagrams 171

6-3 THE ELEMENTS OF BODE DIAGRAMS

So far, we have managed to plot the Bode diagram of an expression having the
form, 1/(1 + jcor). As it turns out, there are other commonly encountered forms of
transfer functions, and we want to know how to plot these also. We refer to these
various forms as elements. Earlier, we discussed another element form, the con¬
stant. We found that when plotted, a constant makes a horizontal straight line for
its gain plot, and its phase angle plot was another straight line that occurred at 0°
regardless of the value of the constant (refer to Figure 6-4).
We now look at yet a third element type. This element is of the form:

Ktf= 1 + jcoT (6-5)


Let’s look at an example:

EXAMPLE 6-3 -
Make a Bode plot of the function, 1 + j'coO.l.

Solution:
Noting that we are no longer plotting a fraction, but instead a whole number expres¬
sion, we again construct a table similar to Table 6-1, Table 6-2.

TABLE 6-2 RESULTS OF PLOTTING THE FUNCTION: 1 + jtuO.l

<0 1 + jcoO.l dBA = 201og(A)

rad/sec result, exp. A E0 dBA


0
0
vq

0.1 1 + 70.01 1.000 0.00


uo
©
O

0.5 1 + 1.001 2.9° 0.01


1 1 + 7'OT 1.005 5.7° + .04
2 1 + 7 0.2 1.020 11.3° + .17
5 1 + 7 0.5 1.118 26.6° + .97
8 1 + 7 0.8 1.281 38.7° + 2.15
10 1 + n 1.414 45.0° + 3.01
15 1 + 71-5 1.803 56.3° + 5.12
20 1 + 72 2.236 63.4° +6.99
30 1 + 73 3.162 71.6° + 10.00
40 1 + 74 4.123 76.0° + 12.30
50 1 + 75 5.099 78.7° + 14.20
80 1 + 78 8.062 82.9° + 18.10
100 1 + 710 10.050 84.3° + 20.00
200 1 + 7'20 20.020 87.1° +26.00
400 1 + 7'40 40.010 88.6° + 32.00
600 1 + 7 60 60.010 89.0° + 35.60
800 1 + 7 80 80.010 89.3° + 38.10
1,000 1 + 7'100 100.000 89.4° +40.00
172 Chapter 6 Bode Diagrams

Plotting this type of expression follows the same analysis as was used for
1/(1 + jcor). The results of plotting individual points from a table and of using the
straight-line approximation method are shown in Figure 6-10. Notice that our
curves are simply upside-down and reversed from what you would have normally
expected after having plotted a fractional-type transfer function. All the phase

ID
U
n
100
Ld
75 l
LD
50 X
<C
25
Ld
0 <X>
<C
X
Cl

,1 1 10 100 1000
(Rad/sec)
Figure 6-10 Bode plot of the function 1 + j'ojO.I.

angles are positive now, as are all the gain values. The slope of the gain curve is
+ 20 dBA/decade to the right of the corner frequency point instead of -20 dBA/
decade. The straight-line curves are drawn from the corner frequency point just
as they were done in the previous example, to form the gain curve. And the phase
angle curve is constructed similarly. That is, the curve is flat ± one decade either
side of the corner frequency.
A fourth type of transfer function element is of the form:

Ktf = C/co)'1 (6-6)

In this expression the value of n can be any positive or negative integer. The
Bode plots for a few positive n values are seen in Figure 6-11. Figure 6-12 shows
the Bode plots for some negative values of n.
After studying these curves for a few minutes, we can see an interesting
correlation between all these curves. To begin with, let’s look at the gain curves
in Figure 6-11. We see that for every singular increase in the value of n, the slope
PHASE ANGLE (DEG)
GAIN (dBA)

Figure 6-11 Bode plots for the transfer function (yco)".

PHASE ANGLE (DEG)


GAIN (dBA)

Figure 6-12 Bode plots for the transfer function \/(j(o)n.

173
174 Chapter 6 Bode Diagrams

of the gain curve increases in an increment of +20 dBA/decade. In addition, the


phase angle curve increases by a +90° increment for each of these changes in n.
Looking at Figure 6-12, we see that as n is incremented, the negative slope of the
gain curve is incremented similarly by an additional 20 dBA/decade. The phase
angle curve is incremented by an additional -90° for each change of n.
Let’s now take a look at yet another element type found in transfer func¬
tions. This is of the form:
1
(6-7)
(1 + jcor)n

Figure 6-13 shows the results of plotting expressions of the form expressed in Eq.
6-7 for several values of n. Notice the similarities between plotting these expres¬
sions and plotting the expressions of (jco)n.
We can conclude from Figure 6-13 that for every increase by 1 in the value of
n, beginning with 1, the negative slope of the gain curve to the right of the corner
frequency increases in increments of 20 dBA per decade. This is similar to the
observations we made for the (jco)n function. We can also see that the phase angle
curve increments by -90° for each increase of n. Notice also the labeling of the x
axis for this graph. Since the corner frequency, wc, always occurs at 1/r, it stands
to reason that for each decade increase or decrease to either side of 1/r, you would
obtain the values of co shown.

0
ID
-90 Ld
(=1
'w'

-180
■270 y
LD
•3605
c
Ld
00
<L
X
CL

.01/T .1/T 1/T 10/T 100/T

wCRad/Sec)
Figure 6-13 Plotting the function 1/(1 -I- jcor)n.
Section 6-5 Graphing Complex Functions 175

6-4 HOW ACCURATE ARE THE STRAIGHT-LINE


APPROXIMA TIONS?

Just how accurate are the straight-line approximation curves when compared to
the true curves? Table 6-3 shows the relative comparison of gain plot values
depending on how close or how far away you are from ojc, the corner frequency.
In other words, these are the corrections to apply to the function 1/(1 + jcor).
If corrections are to be applied to the straight-line approximation curves for
1 + jcor, add instead of subtract the values listed in Table 6-3.
Table 6-4 shows the relative comparison of the phase angle curves for the
straight-line approximation method versus the true curve plot, again depending on
how far or how close you are from ojc .

TABLE 6-3 GAIN CURVE TABLE 6-4 PHASE ANGLE


CORRECTIONS FOR CURVE CORRECTIONS FOR
1/(1 + jcOT) 1/(1 I j(OT)
- -

Corrections to subtract from the Corrections to add to the straight-line


straight-line approximation curve approximation curve to obtain the
to obtain the true dBA gain curve: true phase angle curve:
CO Correction Cl) Correction

0.2a)c 0 dB 0.05a)c. -3'


0.5a>c 1 dB 0.10a>c -6
0.7ojc 1.5 dB 0.15a>f 0'
1.0a»c 3 dB 0.20a>c + 2'
1.3a>c 2 dB 0.50a>c +5'
©
o

1,5coc 1.5 dB 0'


3

2.0oc 1 dB 3.00a)c -5'


3.0coc 0 dB 5.00cac -2'
6.00a)c. 0‘
10.00a>c +6'
20.00a>c + 3'

Again, if you are working with the function 1 + jcor, merely reverse the signs
associated with the correction values listed in Table 6-4.
As you go over Tables 6-3 and 6-4, refer back to Figures 6-8 and 6-10 to
verify these corrections in your mind. It will help you to better understand where
they came from.

6-5 GRAPHING COMPLEX FUNCTIONS

We have already looked at the characteristics of the four elements of transfer


functions. Again, these elements are:
176 Chapter 6 Bode Diagrams

1. Ktf = a constant;
2. Ktf = (>)", (where n is a positive or negative integer);
3. K,f = l/(> + r)n, (where n is a positive integer); and
4. Ktf = 1 + jcor.

Unfortunately, most transfer functions don’t occur in just these simple ele¬
mentary forms. Instead, they are found in more complex forms. That is, they
occur in forms that are in combinations of these four elements. An example of a
complex function would look something like:

0.09
(»2( 1 +>0.2)

We now have to discuss a method for making the Bode plot for this more
complicated and more usual form of transfer function.
Let’s use the preceding transfer function example to demonstrate a method
for constructing its Bode diagram or plot. The first thing we should notice is that
the preceding expression can be broken down into our smaller, less complex,
elements. In other words, we could write the expression as:

0.09 1 1
(6-8)
1 + >0.2
It now becomes obvious that we are dealing with three different elementary
forms of the transfer function: the constant K, (>)", and 1/(1 + jcor). It’s a
relatively simple matter to construct the Bode plot of each of these elements by
themselves. Notice that the final transfer function is nothing more than its indi¬
vidual elements all multiplied together. Since we are plotting the logarithms of the
expressions, we can merely add all the individual curves together to obtain the
equivalent final curve since, in reality, this is the same thing as multiplying all of
these expressions together.
In order to plot Eq. (6-8), you must first plot the individual elements that
make up the expression. This has been done in Figure 6-14. The next step is to
pick several values of a) and begin adding the values of the individual gain amounts
for each curve. Also, add the individual phase angle amounts for each curve at
each co. Pick oo values that are prominently located, such as at the ends of the
curves, at the corner frequencies, and at either side of the corner frequencies.
Table 6-5 shows the results of doing just that for Eq. (6-7). The individual element
plots have been numbered in Figure 6-14 so that you can key the curves to the
results in Table 6-5.
The final result of adding together the individual element curves using the
data in Table 6-5 is shown in Figure 6-14. The final gain and phase angle curves
are shown in boldface type in this figure.
Let’s now try another example of graphing a complex transfer function
expression.
Section 6-5 Graphing Complex Functions 177

PHASE ANGLE (DEG)


TABLE 6-5 DETERMINING THE POINTS NEEDED FOR PLOTTING THE
EXPRESSION 0.09/[>j)2(1 + j(o0.2)]

Gain Phase angle

Curve no.: 1 2 3 Result 4 5 6 Result


00
O
o

+0 +40 = +19 dBA -180 +0 +0


II

0.1 -22
l

CO =

co — 0.5 -22 +0 + 13 = -9 -180 +0 +0 = -180


co — 1 -22 +0 +0 = -21 -180 -15 +0 = -195
co = 5 -22 +0 -26 = -48 -180 -45 +0 = -225
'sO

■'3'
o

10 -22 -6 -180 -60 +0 = -240


II

co =
1
1

co = 50 -22 -20 -68 = -110 -180 -90 +0 = -270


CO = 100 -22 -25 -80 = -127 -180 -90 +0 = -270

EXAMPLE 6-4
Plot the expression:

100
jco{ 1 + jco0.25)

using the straight-line approximation method.


178 Chapter 6 Bode Diagrams

Solution:
Refer to Figure 6-15 for the graph of this expression. We have broken down the
given expression into its elements of 100, 1 /jaj, and 1/(1 + joj0.25). Next, the individ¬
ual elements are plotted just as we did in the previous example. We do this so that
we can keep track of adding the elements together in the proper sequence to give us
our final gain and phase angle summation curves. These final curves are again shown
in bold on our graph.

Figure 6-15 Bode plot for the complex function 100/yoj(I ~h JtoO.25).

6-6 DETERMINING A SYSTEM’S STABILITY

Up to this point, we have learned a system of mechanically plotting the frequency


response curves for transfer functions of devices or systems. Little thought has
been given as to what information is available from these curves concerning the
behavioral characteristics of the system or of the device itself. As we find out
here, we can get a surprising amount of information from our transfer function and
its Bode diagram concerning the behavioral characteristics of the system itself.
As just mentioned, Bode diagrams can be useful in telling us about certain
control system characteristics. Back in Chapter 2 we discussed the characteris¬
tics of positive and negative feedback circuits. In Section 2-6 we discussed how
negative feedback could be used to automatically control the altitude of a hot air
balloon. We described the behavior of an altitude control system that was too
quick, or jittery, in its response, and we also described how that same system
Section 6-6 Determining a System’s Stability 179

would behave if it were too sluggish. What we didn’t mention, however, was the
fact that even though we were inverting the feedback signal by 180° before adding
it to the incoming input command signal, the error signal’s resultant phase angle
(remember, the error signal was the resultant correction or readjustment signal
going into the control system) would be equal to the command signal’s phase angle
(whatever that happened to be) plus -180°. To clarify a point here, let’s look at
our hot air balloon control system once again. Assume for the moment that the
output phase angle, as determined from a Bode plot made of the balloon’s control
system, showed a value of 0°. Remember, the Bode plot is based solely on open-
loop data. Let’s further assume that the input signal had a phase shift also equal
to 0° and a magnitude voltage of, say, 10 volts. In other words, our open-loop
system gain, as determined by the overall system’s transfer function, is:

output 10 Z 0°
system’s transfer function gain, KG
input 10 Z 0°
1 Z 0°

Now, we close the loop on the balloon’s system so that the 180° phase shift takes
place on the output signal in the feedback circuit. Back in Chapter 2 we learned
that the transfer function for a negative feedback system was AB/( 1 + AB). Let’s
assume that we are feeding back 100% of our output signal back to the input so
that now B = 1. Since our transfer function gain is KG, we can now say that the
closed-loop gain of our system can now be expressed as:

KG
(6-9)
1 + KG

We now substitute the gain 1 Z 0° for KG in Eq. (6-9) to see what sort of closed-
loop gain we will obtain:

KG = 1 Z 0° 1 Z 0° 1 Z 0°
= 0.5 Z 0°
1 + (1 )KG 1 + 1 Z 0° 1 + (1 +70) 2
Remember, B = 1

In other words, the correction signal would be reduced by one-half and the
system would undoubtedly respond very sluggishly to this reduced signal. How¬
ever, in all fairness to the preceding problem, we could have selected an output
signal whose magnitude was something greater than 1 volt so that the resultant
closed-loop signal was larger. But, as is typically done in control system design,
system behavior is usually analyzed for a system gain equal to 1, or, as it is more
frequently referred to, the system is usually analyzed at unity gain. This means
for our control system, the output signal would be exactly the same magnitude as
the input signal.
Now, let’s analyze the same system but with an output signal having a phase
angle of —180°, as measured with our open-loop Bode plot instead of 0° just used.
180 Chapter 6 Bode Diagrams

Again, let’s run this new KG figure through Eq. (6-9) to see what the resultant
closed-loop gain will be:

KG 1 L -180° 1 A -180° _ 1 L -180°


1 + BKG ~ 1 + (1)1 L -180° “ 1 + (-1 +j0) ~ 0

Based on what we just calculated, our control system will be jolted with an
infinitely high gain. What started out as an open-loop gain of unity has now turned
into an infinitely large gain, and our control system has now become a wild,
uncontrollable system. This is just the opposite condition of the previous ex¬
periment.
Our two control system examples have pointed out one thing to us. There
must be some compromise existing between an output phase angle of 0° and -180°
where we can expect a control system to behave just right, and there is. We can
find this just-right point of operation either through trial-and-error, or we can use
our Bode plot to obtain this vital information. But before we see how the Bode
plot is used, let’s try another series of experiments having to do with our balloon’s
control system stability. Again, keeping our open-loop system’s gain magnitude
ratio at unity (i.e., this is the same thing as saying a gain of 0 dB), but varying the
output phase angle through the entire range of 0° through 360°, let’s calculate the
resultant gains. Figure 6-16 shows the results. We can see what happens at or

i-
ZD
CL

(—
ID
CL
I-
:d

c
u
CL


_l
I
(=1
u
(y)

_i
u

Figure 6-16 Curve showing relationship between closed-loop gain and the open-
loop phase angle.
Section 6-6 Determining a System’s Stability 181

near -180°; obviously we want to stay clear of that area of phase angle values.
We also want to stay away from the area where the phase angles are near zero or a
little greater in absolute value. It appears that maybe a just-right condition may
occur somewhere in the vicinity where the closed-loop gain is about 1, or maybe
even slightly larger than 1. We don’t want the gain to be too large though, other¬
wise our system may be too quick or jittery. Perhaps we should not go above a
gain of 2 or so. According to our graph in Figure 6-16, this represents a phase
angle range of around -110° to, say, -150°. (This is the same thing as saying a
range of 210° to 250°.) However, this is not a very scientific method of determining
what the desirable range of phase angle operation should be. As it turns out, if the
phase angle is found to be within about 40° to 60° of -180° (see Figure 6-17), the

90

system is going to be stable (i.e., just right) when the system’s control loop is
closed. In other words, a phase angle range of (—180° + 40°) = -140° to (—180° +
60°) '= -120° is the desired design operating range. There is a special name given
to this range; it’s called the phase margin. You might think of it as a margin of
safety. Any phase angles greater than -140° (that is, more negative than -140°)
will result in a too lively or jittery of a system, while phase angles less than -120°
(less negative than -120°) will result in too sluggish of a system. Remember
182 Chapter 6 Bode Diagrams

though, the system’s gain must be 1 (or to say it another way, the system’s gain
must be zero dBA, since 201og 1 = 0 dBA). By the way, if you are wondering how
the figures of 40° to 60° were determined, their derivation is somewhat involved.
At this point we have to accept their values as fact until we learn some more
theory on control systems. Further explanation is given in later chapters.
Notice that in our discussion of phase margin, we referred to that portion of
our curve in Figure 6-16 between 0° and -180° and not the portion from 0° to
+ 180°. The reason for this is partially due to custom. You will find that most
Bode diagrams dealing with negative feedback systems show their phase angles in
the 0° to -360° range rather than in the positive angle range.
Describing the behavioral characteristics of a control system can be a little
difficult. To explain what we mean when we say that a system is either too jittery
or too sluggish is somewhat of a judgment call in certain instances. However,
there are cases where most people would agree that a system is behaving rather
poorly. As an example, think of the steering characteristics of a car. We have, at
one time or another, experienced a car’s steering that was probably very quick—
perhaps too quick for our comfort. In other words, when turning the car’s steer¬
ing wheel rapidly back and forth, we developed a feeling that we were going to
lose control of the car because of the too sudden response or quickness of the
steering mechanism. The other extreme of poor steering is where the car’s wheels
seemed very slow or sluggish in responding to the turning of the steering wheel.
It’s almost as if there were too much slack in the steering mechanism and it took
uncomfortably long for the wheels to respond to the steering command. You had
the feeling that you were trying to turn the steering wheel one way while the car's
body was still moving in the other direction. The steering command and the car's
body were out of phase with each other. Obviously, this was an extreme steering
condition, just as the too quick steering condition was also an extreme. We would
want our steering system characteristics to lie somewhere in between. We have
the similar problem in control system design. We want to be able to design the
just-right system—one that is not too jittery in its response, but one that is not too
sluggish either.
Bode plots are used to forecast a system’s stability habits. They allow you
to make a what-if determination. They allow you to see what would happen if you
were to close the loop on your control system to see how it would behave from a
stability standpoint. Therefore, in reality, Bode plots are open-loop diagrams
based on open-loop information used to forecast closed-loop characteristics.
Based on what we have discussed so far, let’s try our hand at finding out the
stability characteristics of some systems.

EXAMPLE 6-5 --
Using the Bode diagram of Figure 6-14, find the stability characteristics of the plotted
transfer function.
Section 6-6 Determining a System’s Stability 183

Solution:
Step 1. Determine the gain curve’s gain crossover frequency. This is where the
gain curve crosses the 0 dBA line in the Bode plot: According to Figure 6-14, the
frequency at which the gain curve crosses the 0 dBA line is 0.3 rad/sec.

Step 2. At this crossover frequency, determine the phase angle from the phase
angle curve: From Figure 6-14, the phase angle at oj = 0.3 rad/sec is -180°.

Step 3. Determine the phase margin. This is done by performing the following
calculation: phase margin = phase angle reading + 180°. (Note that this is the same
thing as saying: phase margin = phase angle reading — (-180°)).

phase margin = -180° + 180°


= 0°

According to our previous discussion, a phase margin of 0° produces an unstable


system behavior.

EXAMPLE 6-6 --
Find the stability characteristics of the transfer function in Figure 6-15.

Solution:
Step 1. Determine the gain crossover frequency: From Figure 6-15 we see that the
gain crossover frequency is 20 rad/sec.

Step 2. Find the corresponding phase angle at this frequency: Again from Figure
6-15, we see that the phase angle at oj = 20 rad/sec is -165°.

Step 3. Calculate the phase margin:

phase margin = phase angle reading + 180°


= -165° + 180°
= 15°

A phase margin of 15° represents a system that is too jittery since it is outside the
recommended 40° to 60° range and is closer to -180°.
Other examples of determining control system characteristics would be:

1. If, at the gain crossover frequency on a Bode plot the phase angle value were
-100°, the phase margin would be -100° + 180° = 80°. This system would be
too dead or sluggish.
2. On another Bode plot, if at the gain crossover frequency the phase angle were
found to be -135°, the phase margin would be -135° + 180° = 45°. This
system would be quite stable since its phase margin falls within the recom¬
mended range of 40° to 60°.
184 Chapter 6 Bode Diagrams

SUMMARY

In Chapter 6 we discovered that by plotting the transfer functions in a special


configuration called a Bode diagram, we could obtain a much better understanding
as to how the control system whose transfer function we are plotting will behave.
Most transfer functions can be broken down into much simpler expressions called
elements. Each of these elements possesses a particular Bode plot form, so that
by adding these elements together during the course of making a Bode plot, more
complex control system transfer functions may be plotted and analyzed.
The corner frequency on a Bode gain plot is the frequency at which a
relatively sudden change in the gain curve’s slope occurs. This is also called the
gain curve’s break point. Knowing the corner frequencies of a curve helps in the
effort to make an estimate of the shape of that curve’s Bode plot.
The behavior of a control system depends on the amount of gain in the
system and on the resultant phase angle. If we represented a transfer function by
its open-loop form, KG, then its closed-loop equivalent would be (assuming 100%
feedback), KG/( 1 + KG).
Normally, the Bode plot is analyzed for its open-loop phase angle at unity
crossover gain, and then its phase margin determined. The phase margin is the
arithmetic difference between the phase angle and -180°. It is this figure that
gives us insight into the behavioral patterns of the control system. If the phase
margin is less than 40°, the system will most likely be too jittery. If the phase
margin is greater than 60°, the system will most likely be sluggish in its perfor¬
mance. A phase margin between 40° and 60° is considered stable and desirable.

EXERCISES

The following exercises require the using of four-cycle semilog graph paper.
6-1. Draw the open-loop Bode plots of the transfer functions 1, 10, 17, and 125
all on the same graph paper.
6-2. Draw the open-loop Bode plots of joo, (jco)~2, (jco)3, and (joo)4 all on the same
graph paper.
6-3. Draw the open-loop Bode plots of (joo)~\ (joo)2, (joo)~3, and (joo)~4 all on the
same graph paper.
6-4. Plot the expression 1/1 + joo0.05 using the straight-line approximation
method.
6-5. Draw the Bode diagram for the function 15/1 + y’wO.15. Determine its cor¬
ner frequency (in rad/sec), the phase margin, and determine the system’s
stability.
6-6. Draw the Bode diagram for the expression 1 + joo/\ + joo\0.
6-7. Draw the Bode diagram for the expression 21.5/yco(l + jco0.5).
References 185

6-8. Determine the stability of a system whose transfer function is 0.26/(jw)2(l +


jco0.5).
6-9. Do the same for the function 25.5//co(l + j(o0.25).
6-10. Do the same for the function 3(1 + jw0.5)//co(l + jcoO. 1)(1 + jco5). Be
careful with this one as far as plotting accuracy is concerned, due to the
many elements involved.

REFERENCES

DeRoy, Benjamin E., Automatic Control Theory, New York, N.Y.: John Wiley
& Sons, Inc., 1966.
McDonald, A. C. and Lowe, H., Feedback and Control Systems, Reston Va.:
Reston Publishing Company, Inc., 1981.
Sante, Daniel P., Automatic Control System Technology, Englewood Cliffs,
N.J.: Prentice-Hall, Inc., 1980.
Weston Components Division, Servo Engineer’s Handbook, Weston Instru¬
ments, Inc., 1968.
7-1 CLOSING THE LOOP

We found out in Chapter 6 that the Bode diagram was a frequency response curve
especially designed for interpreting the transfer function characteristics of an
automatic control system. All of the data used for constructing the Bode plot
was open-loop data. We used the Bode plot to make a “what if we closed the
loop . . . ?” determination for us. The advantage to this method is that it is fast
and fairly accurate. The disadvantage is, you don’t really see the actual closed-
loop gain and phase angle curves resulting from closing the loop. Perhaps you
want to know what the actual gain and phase angle values are for a given fre¬
quency or range of frequencies. The open-loop Bode plots won’t give you that
information at all. What we need then is some sort of equivalent method that will
allow us to plot a closed-loop Bode plot which will give us this data.

7-2 THE CLOSED-LOOP GAIN PLOT

Since automatic control systems are closed-loop systems, the logical question
arises, are there such things as closed-loop Bode plots? The answer to this ques¬
tion is, yes. However, their construction can be rather difficult because of the
math involved. First, it helps to review the meaning of the general closed-loop
expression, KGH/( 1 + KGH). Look at Figure 7-1. Here we can see the differ¬
ence between the open-loop versus closed-loop system. The transfer function
expressions (i.e., the algebraic expressions used to describe output/input) will be

186
Section 7-2 The Closed-Loop Gain Plot 187

r 1
i 1
I 1
Input- | W KG ■>- Output
1 i
1
1 I
l_ - j

Output , _
-= KG for an Open-loop System
Input

Input

Output KG
for a Closed-loop System
Figure 7-1Open-loop and closed-loop
Input 1 + KGH system comparison.

radically different in each case. Let’s try an example. Assume that we have an
open-loop transfer function for a control system equal to 100/[>(1 + >0.25)].
We have to find its closed-loop output/input expression. The components that
would generate this type of transfer function could be an amplifier (with a magni¬
tude gain setting of 1 used as a summing point for the input control voltage or set
point, and the feedback signal), another amplifier with a magnitude gain setting of
1,000, a servomotor whose transfer function is 5/[>(l + >0.25)] (see Section
4-11.1), and a gear train for reducing the servo’s high rotary displacement output.
The gear ratio for this train is 50 (remember that Kgear = UN; therefore, its
transfer function becomes 1/50). When all of these individual component transfer
functions are combined into one overall system transfer function (this is done by
multiplying all the transfer functions by one another), we will obtain the foregoing
overall transfer function. Actually, there are an infinite number of component
transfer functions that, when properly combined, will give us our desired resultant
transfer function. However, for the purpose of presenting this problem, we use
the combination just given. This system, by the way, could be a system used for
moving a pointer or other indicator to a specific location on a readout scale. The
control signal could be sent from a remote location. The indicator will have the
capability to correct its position automatically if it should be bumped or otherwise
disturbed. Figure 7-2 shows the block diagram for our system. If the servomotor
is forced to move away from a certain position, the servo potentiometer compen¬
sates for this shift in position by outputting a voltage of the opposite polarity,
which is fed back to the differential amplifier (the summing point). A servo pot is
nothing more than a precision potentiometer whose wiper is attached to a rotat-
188 Chapter 7 Closing the Loop

Input Control Voltage

Rotary
Displacement
Output
9

Figure 7-2 Complete automatic control positioning system.

able shaft. The shaft is usually coupled to a servomotor’s shaft. As the servo's
shaft rotates, the potentiometer shaft also rotates, producing a proportional vari¬
able resistance. A voltage is applied to the potentiometer so that a variable output
voltage is obtained that is proportional to the amount of rotation. In our case, this
output voltage’s polarity will cancel a portion of the input control voltage to the
amplifier, thus causing the servomotor to rotate in the opposite direction until it
once again returns to its original position. For our system we will assume 100%
feedback so that H = 1. The conversion to a closed-loop transfer function would
proceed something like this:

100 100
KG _ >(1 + >0.25) >(1 + >0.25)
1 + KG ~ 100 ” >(1 + >0.25) + 100
jco{ 1 + >0.25) >(1 T >0.25)
__100___100_

>>( 1 + >0.25) + 100 7“clt0.25 + joj + 100


100
(7-1)
“ 100 + > - w20.25

As you can see, there is a considerable difference between the open-loop and
closed-loop expressions. Notice that we didn't include a transfer function for the
servo pot in our closed-loop function. The servo pot is part of the feedback loop
and really doesn’t enter into the system’s function. It merely converts displace¬
ment into a voltage. Any device located in a feedback loop generally doesn’t get
included in a transfer function if all it does is convert a quantity into a voltage for
the summing junctions of a system. This conversion must be 100% fed back to the
Section 7-2 The Closed-Loop Gain Plot 189

summing point. Any amount less than this must then show up as a decimal
portion of H in the equation, KGH/( 1 + KGH).
Unfortunately, there is no easy straight-line approximation method that we
can use to plot the closed-loop curves as we had for the open-loop expressions.
The only way to plot the closed-loop expression is by using the laborious point-by-
point calculation method, that is, choose values for to and calculate the resultant
gain and phase angles. (Later, we learn another, much quicker method of con¬
structing these curves.) This can be extremely time-consuming; however, Figure
7-3 is the resultant plot.

ID
UJ

LU
_J
ID
Z
c
Ld
OO
<C
X
CL

After inspecting the curves in Figure 7-3, we can see immediately what one
advantage there is to a closed-loop plot of this kind. We have an immediate
picture of what the gain and phase angle values are for any given value of oj. We
also can’t help but notice the rather prominent feature of the peak occurring in the
gain curve at around 20 rad/sec. This is an often-encountered feature in many
closed-loop gain curves and is called the maximum system gain, or Mm. The
frequency at which the maximum system gain occurs (in our case, 20 rad/sec) is
called the maximum gain frequency, com. And there is an additional piece of vital
information that we can obtain from this peak. As long as this peak doesn’t
exceed a value of 3 dB, the system is probably stable or at worse, jittery, depend-
190 Chapter 7 Closing the Loop

ing on the system’s application. Any value higher than 3 dB describes a system
that is either too jittery or unstable. This interesting piece of information ties in
with a statement that was made in Section 6-6 concerning system stability. If you
recall, we discussed a curve in Figure 6-16 in an attempt to explain how the phase
margin was derived. Referring to this figure, if you were to determine the gain
ratio using the curve at the 40° phase margin, you would obtain a value of approxi¬
mately 1.4. Converting this to dBA, we would get 2.9 dBA. These two observa¬
tions, the 40° phase margin gain and the closed-loop peak gain, are obviously
related and are discussed later.
The gain peak in Figure 7-3 occurs at approximately 14 dBA. This tells us
that the control system described by this curve is entirely too jittery. Inspection
of the system’s Bode plot in Figure 6-15 supports our conclusion. The phase
margin at the crossover point in this diagram is only 15°. This is entirely too close
to -180° for proper system operation.
What would happen to our gain curve peak if we plotted a function that we
knew ahead of time was sluggish rather than jittery? Let’s try it. Let’s plot the
function, 1/(1 + jaj - ar0.25). The reason why we know this particular function is
sluggish, is this: Look at the Bode plot for 100/(100 + jco - 0.25or), Figure 6-15 in
the previous chapter. Notice that we have reduced the constant value, 100, to
only 1. In essence, what we have done here was to reduce the gain by a factor of
201og 0.01, or -40 dBA. In reality, we multiplied the transfer function by the
constant, 0.01, which is the same as subtracting 40 dBA from the original gain
curve. Doing this has no affect on the phase angle curve, since constants plot as
0° on all Bode plots. The resultant Bode gain curve is shown in Figure 7-4.
Reducing the gain curve by -40 dBA has increased the phase margin to a value of
70°. Since this is greater than the recommended 40° to 60° phase margin, the
system will have a deadened behavior. This is the same as reducing the gain of the
amplifier in the system, or increasing the gear ratio in the transmission gear box
connecting the servo to the servo pot. The idea is to tame the system, so to speak,
from what it was originally.
The closed-loop plot of our new function is seen in Figure 7-5. We now see
that the maximum gain hump is missing in the gain curve. This indicates that our
system has in fact become deadened as a result of subtracting the 40 dB from the
original curve. Our system is now too dead. Let’s try to liven it a bit by subtract¬
ing only 20 dB from the original curve in Figure 6-15. Figure 7-6 shows the Bode
plot for the transfer function, 10/[/<x>( + yct>0.25)] that would result from subtracting
this 20 dB. Notice the phase margin has now decreased to a value of only 35°.
This will create a jittery control system. The closed-loop response curve in Figure
7-7 proves this out. Notice the return of the maximum gain hump at co = 6 rad/
sec. The peak of this hump occurs at little over 4 dBA. This verifies the jitteri¬
ness of our system, since the peak of our curve exceeds the recommended +3
dBA value mentioned earlier in our discussion.
Finally, let’s adjust our original curve in Figure 6-15 so that it gives us a
perfectly stable system. We do this so that we can see what the closed-loop
Section 7-2 The Closed-Loop Gain Plot 191

PHASE ANGLE (DEG)


0

-90

-180
GAIN (dBA)

,1 10 100 1000
w(Rad/Sec)
Figure 7-4 Bode plot for l//a>(l + 7'(o0.25).

PHASE ANGLE (DEG)


GAIN (dBA)

w(Rad/Sec)

Figure 7-5 Closed-loop gain and phase angle plot for 1/1 + jw - 0.25to2.
192 Chapter 7 Closing the Loop

PHASE ANGLE (DEG)


GAIN (dBA)

Figure 7-6 Bode plot for 10//co(l + joj0.25).

PHASE ANGLE (DEG)


GAIN (dBA)
Section 7-2 The Closed-Loop Gain Plot 193

frequency response curve looks like. Looking at Figure 6-15, it appears that if we
were to lower our original curve by 30 dB, this would give us a phase margin of
50°. This would put us right in the middle of the recommended phase margin
region for stability. Figure 7-8 shows the Bode plot and how the phase angle value
was determined. The resultant closed-loop frequency response curve for this
system is seen in Figure 7-9. The gain curve in this figure shows a barely percepti¬
ble rise in gain before taking the characteristic drop to the right of this rise. The
maximum gain at the peak is less than 0.2 dBA. This very gradual rise in the gain
is fairly characteristic of a stable system although it varies from system to sys¬
tem. Again, this rise may be as high as +3 dB and still produce an acceptable
system.

ID
Ld
n
Ld
_l
LD
Z
<C

Ld
(A
C
X
CL

Figure 7-8 Bode plot for 3.2//co(l + ja)0.25).

Another important piece of information that we can obtain from our closed-
loop curves, in addition to the maximum peak gain and frequency, is the system’s
bandwidth. In Section 2-3 of Chapter 2 we defined bandwidth as the span of
frequencies that is included within the 3-dB down points of a device’s frequency
response curve. Applying that definition to Figure 7-3, we see that the left-hand
side of the gain curve is relatively flat and at approximately 0 dBA. It appears to
be quite unlikely that the curve would go down to -3 dBA before reaching 0 rad/
sec. However, the right-hand side of the gain curve is seen to dip to -3 dBA at 32
rad/sec. Therefore, the bandwidth of this curve is said to be equal to 32 rad/sec.
System bandwidths, such as the one we have been working with here, are always
194 Chapter 7 Closing the Loop

PHASE ANGLE (DEG)


w(Rad/Sec)
Figure 7-9 Closed-loop gain and phase angle plot for 3.2/1 + jco - 0.25oP.

determined from their closed-loop data, never from their open-loop Bode dia¬
grams.
Another observation that we can make concerning the maximum system
gain is its frequency value on the closed-loop response curve. Its frequency al¬
ways coincides with the frequency of the gain crossover point on the Bode dia¬
gram. Be sure to verify this to yourself using the curves we have just finished
discussing. This is a helpful check in comparing the accuracy of plotted data in
both diagrams.
So far, we have focused our attention on the gain curves in our closed-loop
response curves. Let’s look now at the phase angle curves to see what informa¬
tion they contain for us. Earlier, we stated that the transfer function we have been
working with represents the control system in Figure 7-2. Let’s use the modified
version of this function as in Figure 7-9. According to this curve, the bandwidth is
now only 4.3 rad/sec, as opposed to the earlier bandwidth of 32 rad/sec when the
amplifier gain was higher (or the gear box ratio was lower). The actual frequency
at this point is equal to co/2tt, or 0.68 Hz. In other words, if the command voltage
variations on our servomotor vary any more rapidly than about 0.7 Hz, the sys¬
tem, in all likelihood, will not be able to keep up with these variations. And notice
what the phase angle is doing when we are near the maximum gain area of the gain
Section 7-3 Using the Nichols Chart 195

curve. The closed-loop phase angle curve begins changing rather rapidly. This is
especially true for livelier systems such as the one described in Figure 7-7 (or such
as in the extreme case seen in Figure 7-3). Obviously, it isn’t desirable to have a
system’s output out of phase with its input. This is like the car’s steering mecha¬
nism described in Chapter 6. This could cause a lot of confusion for the system’s
mechanisms and electronics, not to mention the amount of confusion created for
the system’s operator. We want to operate in that portion of the closed-loop
frequency response curve where the phase angle remains very close to 0°. This
is usually in the region just to the left of the maximum gain point on our gain
curve.
Let’s review the advantages of using closed-loop data instead of open-loop
data in designing a control system. A closed-loop frequency response plot allows
us to:

1. determine the bandwidth of the control system;


2. find the maximum system gain, Mm;
3. determine the frequency at which the maximum system gain occurs; and
4. find the actual phase angle occurring at a particular frequency.

In the next section we learn another method for constructing the closed-loop
diagram allowing us to obtain the very same information but in a much shorter
time.

7-3 USING THE NICHOLS CHART

As we have found out, constructing closed-loop frequency response curves can be


a very tedious and time-consuming experience. The calculations alone account
for the majority of time spent in preparing these diagrams. It helps to have a
programmable calculator available for performing the numerous math steps re¬
quired; however, we now look at a much easier method that’s available to us.
This is a method that produces fairly accurate and quick results. We will have to
turn in our log paper used for making Bode plots for a special coordinate paper
referred to as Nichols chart paper. A copy of this paper style is seen in Figure
7-10. Unfortunately, this paper is not as readily obtainable as log paper; there¬
fore, it will be necessary to make photocopies of Figure 7-10. Doing this will
allow us to work the problems that we discuss here.
Actually, we will still be using log paper for construction of the open-loop
Bode plot. This information is needed so that we can use the Nichols chart to
convert the open-loop data into closed-loop data. We now analyze this new
method and, in the process, use an example problem to see how the conversion
process works.
Gain in decibels

-180 -160 -140 -120 -100 -80 -60 -40

Phase angle in degrees

Figure 7-10 Nichols chart.

196
Section 7-4 An Example 197

7-4 AN EXAMPLE

Let’s convert the open-loop Bode diagram for the expression 100/>(1 + >0.25)]
into its closed-loop response curve. We will have to refer back to Figure 6-15 for
the open-loop Bode plot of this expression. Here is a step-by-step instruction on
how to use the Nichols chart method.

Step 1: Construct the open-loop Bode plot of your control system as usual
(Figure 6-15).

Step 2: Convert your straight-line approximation curves to smoothed


curves using the conversion information given back in Section 6-4, Chapter 6.
(See Table 6-3 in this section for the gain conversions and Table 6-4 for the phase
angle conversions.) The results of our example conversion are shown in Figure
7-11.

0 CD
Ld
n
-90 u
_l
LD
Z
-180 <C
Ld
(A
C
X
CL

,1 1 10 100 1000
w<Rad/Sec)

Figure 7-11 Smooth-curve Bode plot of 100//a>(l + 0.25jco).

1 Step 3: Choose about 10 or so to values from your Bode plot and determine
both the gain and phase angle values for each of the co values. If necessary, make
a table of these results (see Table 7-1). Obtain a sheet of Nichols chart paper and,
using the chart’s outer margin rectangular coordinates of gain and phase angle,
transfer the preceding gain and phase angle values for each o> to this new grid.
198 Chapter 7 Closing the Loop

TABLE 7-1 OPEN-LOOP GAIN AND


PHASE ANGLE VALUES FROM
FIGURE 6-15

(X) Gain (dBA) Phase Angle (deg)

4.0 25 -135
6.0 20 -147
7.0 17 -150
10.0 11 -159
15.0 4 -166
20.0 -1 -170
30.0 -8 -174
40.0 -13 -174
50.0 -17 -175
70.0 -22 -178

Label each co with its value. With a french curve, draw a smooth curve through
your 10 or so points, as was done in Figure 7-12.

Step 4: At each w point on the gain-phase angle curve, read off the closed-
loop gain values using the curved grid pattern labeled Gain KG/(l + KG). Also,
read off the closed-loop phase angle values using the curved grid pattern labeled
Angle KG/( 1 + KG). Record these new data in a separate table (see Table 7-2).

TABLE 7-2 CONVERTED


CLOSED-LOOP DATA TAKEN FROM
NICHOLS CHART

c0 Gain (dBA) Phase Angle (deg)

4.0 .4 -3
6.0 .7 -4
7.0 1.1 -5
10.0 2.7 -7
15.0 8 -20
20.0 9+ -123
30.0 -3.5 -170
40.0 -11 -172
50.0 -16 -174
70.0 -22 -178

Step 5: Plot the data obtained from the Nichols chart onto semilog paper.
The results of our example conversion are shown in Figure 7-13.
25

20

15

10

-5

10

-15

20

25
-1 -160 -140 -120 -100 -80 -60 -40

Phase angle in degrees

Figure 7-12 Using the Nichols chart for plotting Example 6-6.

199
200 Chapter 7 Closing the Loop

0
-20
S~\

ID
-40 Ld
-60 nV

-80 u
_J
-100 LD
X
-120 <r
u
-140 </>
<c
X
CL

.1 10 100 1000

w(Rad/Sec)
Figure 7-13 Closed-loop plot generated from using the Nichols chart method (see
Figure 7-3).

As you can see, the results in Figure 7-13 compare fairly well with the hand-
plotted results in Figure 7-3. There is some distortion present in the curves of
Figure 7-13; however, this is not a serious problem as far as interpreting the data
in the resultant curves. The maximum peak gain value obtained with the Nichols
chart does not go any higher than 9 dBA, but since this is an unstable condition to
begin with, there is really no need to go beyond this reading. Since our Nichols
chart doesn’t convert open-loop data beyond +25 dBA or below -25 dBA, the
resultant closed-loop curves will be cut off at these points. Again, the trend of
most curves at these points is generally established, so you can easily extend the
curves, if you like, with little loss in accuracy. In our case, the curves’ ends are
obviously headed outward along straight lines and may be extended as straight
lines if desired.
The most interesting thing about using the Nichols chart method has to do
with its conversion speed. Being able to convert rapidly from the open-loop data
of a Bode plot to the closed-loop data using the Nichols chart, one can easily
observe the closed-loop frequency response curves as they respond to any modifi¬
cations being made to the transfer function. The problem we have been working
with is a good example. We know, for instance, that the transfer function, 100/
Section 7-4 An Example 201

[>( i + ja)0.25)] is unstable. We ascertained this when we plotted its Bode dia¬
gram in Figure 6-15, only to discover the very small phase margin. We verified its
instability when we hand-plotted its closed-loop transfer function and saw the
excessively large maximum gain hump that exceeded the +3 dBA criteria. This
fact was proven in the Nichols plot. However, what we could not ascertain from
our frequency response curve was how to correct the unstable condition. Some
trial-and-error in adjusting the values of our transfer function was necessary in
order to fine-tune the correct values needed in the function. With the Nichols
chart, we can now keep an eye on the gain-phase angle curve to see if we are
compensating our transfer function by the correct amount necessary to create the
desired results in the closed-loop plot.
Earlier, we found that by lowering the system gain by 30 dB, we could obtain
a desirable phase margin of 50°. We determined this by inspecting the open-loop
Bode diagram for our transfer function (Figure 6-15) and noting that 30 dB was
needed in order to create a desired 50° phase margin. This 30 dB reduction cre¬
ated an open-loop transfer function of 3.2/[/w(l + 0.25jo>)]. Let’s plot this cor¬
rected transfer function using our Nichols chart to see if we obtain the same
results as were obtained earlier plotting it by hand. We pick off our open-loop
gain and phase angle values from the curves already constructed in Figure 7-8.
The resultant Nichols gain-phase angle curve is seen plotted in Figure 7-14. The
final closed-loop gain and phase angle curves are illustrated in Figure 7-15. Our
results compare favorably with the hand-plotted results of Figure 7-9 with only
minor discrepancies. Notice, however, the bandpass results. We obtained a
bandpass of 5 rad/sec using our Nichols method, whereas the hand-plotted point-
by-point method showed a bandpass of 4 rad/sec. This points out the possibility
of discrepancies that can occur between these two methods. Only carefully laying
out points on paper and carefully reading the data on the Nichols chart can help to
reduce these differences.
There is yet a third method for obtaining closed-loop response curves for
control systems, and it is perhaps the most accurate and quickest method of all.
The method involves using a personal computer to perform all calculations and to
automatically plot all data results. Software is available which can be modified to
perform all open or closed frequency response plotting. The operator only has to
enter the design parameters according to prompts given by a screen menu and the
data is plotted in a very short period of time. Again, because of the speed of
plotting by a dot-matrix printer, the operator can make modifications to the sys¬
tem and can see within seconds how the overall system will respond to these
modifications. We should now ask ourselves this question: Of the three methods
that we have now studied, the manual point-by-point method, the Nichols chart
method, and the computer method, which is the most accurate? We would
have to state, of course, the computer method. Not only is it unquestionably the
most accurate method, but it is also the fastest method of obtaining plotted re¬
sults.
Gain in decibels

-180 -160 -140 -120 -100 -80 60 -40

Phase angle in degrees

Figure 7-14 Plotting 3.2//co(l + 0.25/co) using the Nichols chart.

202
Summary 203

0
-20

PHASE ANGLE (DEG)


-40
-60
-80
-100
-120
-140

,1 1 10 100 1000
(Rad/Sec)

Figure 7-15 Closed-loop gain-phase curves for 3.2/1 + jco - 0.25a>2.

SUMMARY

The Bode plot is used for predicting closed-loop behavior. The actual closed-loop
data must be obtained by converting the open-loop transfer function to a closed-
loop function and plotting the results. This can be a rather tedious process. On
the other hand, the Nichols charting method is a fairly convenient method for
generating a closed-loop plot from the open-loop transfer function. The quickest
method, however, is through the usage of a personal computer and the several
software programs that are now available for this sort of application.

EXERCISES

7-1. Plot the closed-loop frequency response curves (i.e., both the gain and
phase angle curves) for the expression:

jaj(\ + jco0.2)

7-2. Plot the closed-loop frequency response curves for the expression:
204 Chapter 7 Closing the Loop

370
KG
XI +>0.25)

What is the stability condition for this system?


7-3. Find the open-loop bandwidth for the system:

XI +>0.2)
Compare this figure to the closed-loop bandwidth for the same system. Do
you see any correlation between the two figures?
7-4. Plot the closed-loop frequency response curves for the expression:

KG __1
1 + KG 1 + jco — 0.25co"

What hints do you see from observing these curves that this system may be
sluggish?
7-5. Construct the Bode plot for the expression:

200
KG
>(1 + 0.016)

Find the phase margin. Determine the gain reduction needed to make this
system stable. Now, using the Nichols chart method, construct the closed-
loop gain-phase angle frequency response curves and determine again the
amount of gain reduction needed to stabilize the system. How close are
your two gain reduction results?
7-6. Construct the closed-loop frequency response curves for the expression:

25.5
KG
jco( 1 + >0.25)

7-7. For the open-loop expression in Exercise 7-1, plot the closed-loop fre¬
quency response curves using the Nichols chart method.
7-8. Find the peak frequency in the gain curve and calculate the gain and phase
angle at this frequency for the transfer function:

200
KG
>( 1 + >0.333)

7-9. Using the Nichols chart method, determine the reduction in gain needed, in
dB, to make the transfer function in Exercise 7-3 stable. Hint: The system
is considered stable when the gain curve’s peak gain is less than +3 dB.
7-10. Refer to Figure 7-3. Determine the change in bandwidth resulting from
reducing the peak gain in the gain curve to make this system stable.
Transient Analysis

8-1 INTRODUCTION

One of the most interesting aspects of automatic control systems has to do with
the system’s behavior when it is first turned on, or when it receives a change in
command signals. Many control system designs that have been scrapped due to
poor performance would have probably been otherwise acceptable if all they had
to do was to operate with very slow or infrequently changing command signals.
Obviously, this is an unrealistic condition for any control system. Control sys¬
tems must be designed to handle rapidly changing input commands and must be
capable of remaining stable during this time. In this chapter we study ways of
forecasting a control system’s behavior to a sudden change in command signal and
how to correct an otherwise undesirable behavioral characteristic. A system’s
reflex to a command signal is a relatively short-lived event; consequently, we refer
to this condition as a transient response. The term transient means that the
system’s response or behavioral reaction to a command signal is only temporary;
it is a passing condition that shortly gives way to a more sustained steady state
condition.

8-2 A SYSTEM S RESPONSE TO A STEP INPUT

In order to analyze a control system’s transient response, we must have some,


means of measuring the control system’s output response over very short periods
of time. We refer to time periods measured in milliseconds, or even in microsec-

205
206 Chapter 8 Transient Analysis

onds. To get a better grasp of what we are describing here, let’s look first at a
system described in Figure 8-1. This is an automatic chart recorder used for the
automatic recording of a variable DC input voltage. This is how it works: A
variable DC voltage is fed into one of the two inputs of a differential op amp (A)
whose gain is 100. From there, the amplified output signal is fed into another

amplifier (servoamplifier B) having an additional gain of 5. The purpose of this


amplifier is to supply the required power to drive the servomotor. (The op amp by
itself does not have the required power needed to do this; it only has voltage gain
capabilities.) The output of the servoamplifier is then fed to the servomotor where
it rotates in either direction, depending on the original polarity of the input DC
voltage. As the motor rotates, a rack and pinion arm moves in and out across the
moving chart paper beneath a pencil attached to the moving rack. Attached to the
pencil is a wire string that stretches back to a pulley and string pot assembly. A
string pot is nothing more than a potentiometer attached to a spring-loaded coil of
string, called a yo-yo. As the spring-loaded string is pulled into and out of the
string pot, a variable resistance, transformed internally to a variable voltage, is
created at an output terminal on the pot. This variable voltage is sent to the
second input of the op amp.
With the switch shown in our system in its off position, we have an open-
loop system. With the switch in its on position, we have a closed-loop system. In
the closed-loop configuration, the output of the string pot is connected to the op
amp. This output is adjusted so that it is exactly the same amplitude as the
Section 8-2 A System’s Response to a Step Input 207

variable input voltage being supplied to the other input of the op amp. We now
forcibly place the pencil over to the edge of the moving chart paper with our
hand. This position will represent 0 volts input into our system. Now we supply a
5-volt DC voltage to our system’s input, causing the servomotor to jump instantly
into action moving the rack and pinion with its pencil over to the other side of the
chart paper. However, the servomotor keeps on moving beyond the paper, since
nothing is telling it to do otherwise. The only way to stop the motor’s turning is to
remove the 5-volt input voltage. This is obviously not a very practical chart
recorder design.
To correct this situation, we again force the pencil back over to the far side
of the chart paper. This time, however, we adjust the output of our string pot,
which is tracking the position of the recording pencil, so that it reads 0 volts.
(Voltage adjustments are made to the string pot by means of an internal pot that
varies the supply voltage to the yo-yo pot.) Next, we forcibly move the pencil
over to the opposite edge of our chart paper (assuming that this is where we want a
5-volt signal to cause the recording pencil to move to, to make its 5-volt recording
mark) and again adjust the string pot, but this time for a 5-volt output reading.
Now, we apply the 5-volt input signal. Again, our servomotor jumps into
action, causing the pencil to move once again across the moving chart paper. This
time the pencil stops at the paper’s opposite edge, exactly where we had manually
positioned the pencil earlier. What has happened is this: As soon as the string pot
outputted 5 volts to the op amp, the op amp’s output dropped to zero since its
output was equal to the difference between its two input voltages (times its gain of
100, which still results in an output of 0). What would happen if we supplied, say,
2.5 volts to the op amp instead of 5 volts? Our pencil recorder would then travel
only as long as there was a difference between the op amp’s two input voltages, or
until the string pot outputted 2.5 volts. This would occur when the pencil was
mid-way across the surface of the chart paper.
Let’s inspect the chart paper results from applying the 2.5-volt input signal
to our voltage recording system. Refer to Figure 8-2. As you look at the chart
paper, notice that time can be measured along its length. Let’s assume that each
division represents 100 milliseconds of time. What we are interested in is the zig¬
zag waveform, called a damped sinusoidal wave, that was produced from time = 0
to about time = 200 ms. This is our transient response condition for this particular
control system. In addition, this is called a transient response to a step input.
The term step input refers to the fact that a sudden 2.5 volts was applied to the
system, not a gradually increasing 2.5 volts. The characteristics of these two
conditions are important. If the 2.5 volts were applied gradually, we would get a
distorted-looking transient response curve, one that would contain the damped
sine wave feature but along a gradually increasing line. Also, the amplitudes of
each individual cycle within the waveform would be reduced. We will see why in
a moment.
The transient waveform is produced by a number of quantities. The cyclic
homing of our recording pencil is called hunting. Hunting is the result of overcor-
208 Chapter 8 Transient Analysis

rection and undercorrection of the system’s response to a command signal.


Ideally, the system’s output (the pencil, in our case), would report to its exact
location with no hunting. The hunting of a control system’s output is caused by
component inertia. In other words, in the case of the moving pencil and the rack’s
steel-machined arm, these components have a difficult time stopping and starting
instantly because of their respective masses. Also, the frequency of oscillation is
determined by several factors, as we soon discover. One such factor has to do
with the system’s frequency response, and another has to do with its time con¬
stants, velocity constants, and the amount of viscous damping built into the servo¬
motor (see Section 4-11.1, Chapter 4).
There are other types of input waveforms that could be supplied to the
automatic control system. Figure 8-3 shows some of these types. However, the
step input is the one most frequently used for the reasons mentioned earlier. This
is a square-wave input rising from zero volts to some higher voltage value, usually
one volt, rising in an infinitely short period of time. More precisely, this type of
input is called a unit step input, and is shown graphically in Figure 8-4. We use the
unit step input to “shock” our control systems so that we can then study their
behavioral responses. This type of input stimulus allows us to study the various
characteristic outputs that will enable us to evaluate the performance of our
systems.
Section 8-2 A System’s Response to a Step Input 209

Figure 8-3 Possible input waveforms


(d) Square Wave for an automatic control system.

Note: Normally, t = 0 Begins at the Axes Intersection, but


is Shown to the Right of Intersection for Clarity Figure 8-4 A unit step input signal.
210 Chapter 8 Transient Analysis

8-3 GENERATING AND ANALYZING A TRANSIENT RESPONSE


CURVE

Before closing the switch in the feedback circuit of our recording device, our
open-loop transfer function was:

d (Ax)(A2)(Kv)(Krp)
Ec 5(1 + st)(N) " j

where d the output displacement of the rack’s arm (in.)


Ec the control voltage supplied to the system (volts)
Krp transfer function for rotary pot (in./rad)
Ky velocity constant of servomotor (rad/sec/volt)
r servomotor’s time constant (sec)
N gear ratio for rack and pinion plus gear train in Figure 8-1

Let’s assume the following values for all the variables in Eq. (8-1):

A\ = 100

A2 = 10

Ky = 10 rad/sec/volt
r = 0.2 sec

N = 100

Krp = 0.25 in./rad


After substituting these values into Eq. (8-1), we have for our open-loop
transfer function,

d_ = (100)(10)(10)(0.25) = 25
Ec 5(1 + 0.25)(100) 5(1 + 0.25) 1 j

Notice that a transfer function for the string pot was not included in the
transfer function, since it had nothing to do with the open-loop functioning of the
system. Even when we close the feedback loop, we still ignore the string pot for
the same reason we ignored the servo pot in our control system back in Chapter
7. The string pot merely converts the positional information coming from the
servomotor into a proportional voltage that is fed back to the system’s summing
junction. As in our previous system, we will feed back 100% of this signal so that
H = 1.
25 25
KG 5(1 + 0.25) 5(1 + 0.25)

1 + KG 25 5(1 + 0.25) + 25
1 +
5(1 + 0.25) 5(1 + 0.25)
Section 8-3 Generating and Analyzing a Transient Response Curve 211

25 25 __25_
5(1 + 0.25) + 25 s + 0.25“ + 25 jco + 0.2(Jol>)~ + 25
25
(8-3)
25 + jco — 0.2 a)2

Equation (8-3) is now our closed-loop transfer function which we can now
plot to see its closed-loop characteristics. We also plot the open-loop Bode plot to
determine the system’s phase margin. Figure 8-5 is the Bode plot and Figure 8-6
shows the closed-loop frequency response curves.

(Rad/sec) PHASE ANGLE (DEG)


Figure 8-5 Bode plot for 25!s(\ + 0.2s).

We can see from the Bode plot that the phase margin for our system is only
28°. This tells us that our system is entirely too jittery to be of any use. We would
find that whenever a command voltage, ec, would be supplied to the chart re¬
corder, the pencil would overshoot the desired voltage mark many times before
eventually settling at the proper voltage mark. Of course, our closed-loop plot
shows us the characteristic gain hump that exceeds the desired 3 dB limit of
stability. This verifies what the Bode plot told us. While we are looking at the
closed-loop plot, let’s make a note of our bandwidth. We see that it is 16.5 rad/
sec, or the equivalent of 2.6 Hz. We refer to Figure 8-6 later.
212 Chapter 8 Transient Analysis

PHASE ANGLE (DEG)


w(Rad/Sec)
Figure 8-6 Frequency response curve for 25/25 + joj - 0.2a>2.

8-3.1 The Natural Resonant Frequency and the


Damping Factor

We want to be able to determine some additional information about our control


system. It would be helpful to be able to obtain a graph of the hunting characteris¬
tics of our marking pencil as it searches for its final resting spot at the proper
voltage mark on the chart paper. If we could do this, then we wouldn’t have to
run the chart recorder to see how it would behave. We could, instead, draw its
transient curve out on paper ahead of time. However, in order to do this, there
are two quantities that we must first calculate. We want to be able to find (a) our
system’s natural resonant frequency, ajn, and (b) the system’s damping factor, z.
The natural frequency of any system is the frequency at which the system oscil¬
lates, rings, or vibrates naturally without any outside influence except for the
initiating energy or force. The damping factor, or damping ratio as it is sometimes
called, is the rate of dying-out of the oscillations. We will have a better under¬
standing of this term in a few moments.
To find ajn and z, we first rewrite our closed-loop function in its Laplace
form: 25/(0.2s2 + 5 + 25). The next thing we do is to supply a unit step voltage
input to our system, but do it mathematically. Referring to the table of Laplace
transforms in Chapter 3 (Table 3-1), we see the Laplace equivalent of a unit step
Section 8-3 Generating and Analyzing a Transient Response Curve 213

input as 1/s. In reality, we would apply a sudden one-volt shock to our system.
However, the mathematical equivalent of doing this is:

1 25
(8-4)
5 0.2s2 + s + 25

Referring again to the Laplace tables, we search for a transform having the de¬
scription of a closed-loop second order system (i.e., a system having terms raised
to the power of 2) being excited by a unit step input of 1/s. In Table 3-1 we find
Eq. 17 that fits the description. However, we first divide the entire expression,
both its numerator and its denominator, through by the term, con2, to give:

1
(8-5)
2zs
s + 1
(On

The reason for doing this will become evident in a moment. Meanwhile, if we can
get Eq. (8-4) to somehow confirm to Eq. (8-5), we can then find the values of con
and z.
To get Eq. (8-4) to look like Eq. (8-5), we first divide the numerator and
denominator by 25, to get:
1
(8-6)
r 0.008s2
s
- 1
Next, we multiply the denominator of Eq. (8-6) by 125/125 to get:

1
(8-7)
125s
+ 1
3,125
Since the first term in the denominator (125) is w2, the second denominator term
(3,125) must be “forced” into being con, which is V125 or 11.2. So we multiply
this second term by 11.2/3,125. Of course, we have to do the same thing to the
125s numerator term. Our expression now becomes:
J_
11.2 0.448s
125s + 1
3,125 11.2 (8-8)
+ + 1
125 3,125 11.2
3,125
Now that Eq. (8-4) is in the form of Eq. (8-5), we can identify con as 11.2 rad/sec
(i.e., a frequency of 1.78 Hz) and z as 0.224 (be sure to note that 2z = 0.448, so you
have to take \ the value shown in Eq. (8-8). A damping factor of 0 means that a
control system will oscillate forever, each peak of oscillation having the same
amplitude. A z of 1 means that there are no oscillations; the output, in other
214 Chapter 8 Transient Analysis

words, reports directly to its final location or value with no hunting whatsoever.
However, the time it takes to get to that spot will be excessively long. This is
called a critically damped system. Consequently, a z of 1 represents a very slug¬
gish system and is as undesirable as a z value near 0.
At this point it is logical to ask, what is considered an acceptable value of z?
From experience, it has been found that control systems exhibiting any values less
than 0.4 tend to be too jittery; any values greater than 0.7 and the systems tend to
be too sluggish. Therefore, any z values within the range of 0.4 to 0.7 are consid¬
ered acceptable by many control system designers.

8-3.2 The Damped Frequency

Our system is obviously too jittery. And while we discuss oscillations here, we
must be able to distinguish the difference between a system’s natural resonant
frequency and that system’s damped frequency. Earlier, we implied that a sys¬
tem’s natural frequency of oscillation was its unimpeded frequency. That is, the
system would oscillate at its natural frequency of oscillation if there were no
losses due to friction, viscous damping, and so on. In reality, since these impe¬
dances are definitely present, the system will instead oscillate at a somewhat
lower frequency called its damped frequency, cod. The quantity ajd is related to con
by the equation:

0>d = 0>n Vl - z2 (8-9)


Figure 8-7 gives us a visual idea as to how the damping ratio values alter the shape
of the transient curve. These are generalized curves for different z values. Note
in particular how the z values produce the varying amounts of a)d.
We are now armed with enough information to allow us to actually plot the
transient waveform of our system to see what sort of settling pattern will be drawn
by the pencil as it eventually settles at a final reading. However, based on our
computed z value of 0.224, our recorder’s pencil may take a few oscillations
before settling down to its final reading.
Let’s now refer to our table of Laplace transforms once again in Chapter 3.
We want to find the time domain equation for Eq. (8-5) because, knowing this, we
can find the amplitudes for any value of time in our transient function. According
to Table 3-1, Eq. 17, the time domain equation equivalent is:

1 + (onV 1 - z2 t - arctan (8-10)

where the arctan expression returns an angle, </>, that must he greater than 0 but
less than 180°.
Equating our preceding expression, Eq. (8-10), to y, and using our values for
z, <*)„, and o)d, we can now hand-plot the damped transient waveform for our
control system. There is one modification, however, we must make to Eq. (8-10)
Output Section 8-3 Generating and Analyzing a Transient Response Curve 215

Figure 8-7 Comparison of damping


factor (z) values.

in order for it to have consistent units. Notice that the expression is taking the
sine of an expression containing ajn. Since con is in radians, we have to remember
to convert the expression to degrees before taking the sine of it. Therefore, we
modify Eq. (8-10) by multiplying the ajn expression by 57.3°/rad to convert it to
degrees. We also equate the expression to y, since this is a relationship that
allows us to determine an amplitude, y, for a given value of time, t. Equation
(8-10) now becomes:

y = l + sin (to„Vl - z2 • f)57.3 - arctan (8-11)

Using rectilinear coordinate graph paper, we can now lay out a y (amplitude)
versus t (time) axis. However, it will take a considerable number of points to
construct our curve, but there are some additional calculations we can do to make
our plotting job easier. First, let’s look at a typical transient waveform. Figure
8-8 shows such a waveform and how our recorder’s waveform will probably look
also. Notice the curve begins at 0, crosses the 1 axis (which is the desired final
resting position for this system, and for our system too) and reaches a peak
216 Chapter 8 Transient Analysis

Figure 8-8 The transient response curve as a result of a unit step input quantity.

amplitude before going back down across the center axis once again. The cycle is
repeated a number of times, each cycle having less of an amplitude as compared to
the previous cycle. This is characteristic of a damped transient waveform. It
would be helpful to us if we could calculate the crossover times, t\, t2, h, U, etc.
Then, we could calculate the maximum and minimum amplitudes occurring mid¬
way between each crossover time and that would give us a pretty good idea what
our final curve is going to look like. We can find the crossover time each time y =
1 in Eq. (8-11); in other words, when

1 + sin (a>„Vl - z2)57.3 - arctan = 1 (8-1 la)

This event will take place whenever the sine term in Eq. (8-11) becomes 0. That
is, when

,_ Vl - 72
(co„Vl - Z2 • /)57.3 - arctan-— = 0 (8-1 lb)
—z

We solve the preceding equation for t, but first, let’s substitute the expression cnd
for the expression ou„V 1 - z2, since they are equivalent (Eq. 8-9). We now have,

(ajd)513t - arctan = 0 (8-1lc)

Solving for t:
Section 8-3 Generating and Analyzing a Transient Response Curve 217

Vl ~ Z2~
J —z
t = ta = arctan-jz-z— (8-1 Id)
cod j/.i

where ta = initial crossover time (sec)

Since one cycle in our transient waveform is cod in length, and one cycle represents
277 radians, and a crossover occurs at co = 2ir!T intervals, in general, we can say
that o)d = 2jt/T, or

T = — (8-1 le)
<°d

where T = the waveform’s period (sec)

But, since there is a crossing every 180° in each cycle, that is to say, a crossing
every half-cycle, we have

T=2tb = — (8-1 If)


(l)d

where tb = a crossover time interval per half-cycle (sec)

or,

(8-1 lg)

These crossovers will occur for n cycles; therefore, we can modify Eq. (8-1 lg)
further by multiplying the expression by n, and letting n become a series of integer
values until the waveform we are plotting dampens out completely. Therefore,
our final expression for determining all the crossover times occurring on the 1 axis
of our transient plot will be the sum of ta (Eq. 8-11) and the n modified tb expres¬
sion, in Eq. (8-1 lg). That is,

arctan(Vl - z2/~z) nzr


^x-over + — (8-12)
57 3o)d <*>d

where n = 0, 1, 2, 3, 4, etc.

For each crossover time determined by Eq. (8-12), we can place the mid-value
between any two crossover times into Eq. (8-11) to find y, the amplitude at that
particular time.
Table 8-1 lists the calculated terms resulting from using Eqs. (8-11) and (8-12)
in developing the transient response curve for Eq. (8-4). The first three amplitude
values listed in Table 8-1 are calculated in the following manner:
The first crossover time, t\, for Eq. (8-4) requires using Eq. (8-12). The
following constant values are obtained from Eq. (8-8):
218 Chapter 8 Transient Analysis

TABLE 8-1 CALCULATED CROSSOVER


TIMES AND AMPLITUDES FOR
TRANSIENT CURVE RESULTING
FROM APPLYING A UNIT STEP INPUT
TO EQ. (8-4)

n t Oc-over Mid-value Amplitude

0 h 0.165 1.000
0.309 1.473
1 h 0.453 0.998
0.597 0.771
2 h 0.741 1.001
0.885 1.111
3 U 1.029 0.999
1.173 0.946
4 h 1.317 1.000
1.461 1.026
5 h 1.605 1.000
1.749 0.987
6 C 1.893 1.000
2.037 1.006
7 h 2.181 1.000
2.325 0.997
8 t9 2.469 1.000

con = 11.2 rad/sec

z = 0.224

ajd = 10.9 rad/sec, using Eq. (8-9)

Since we are interested in the crossover time for n = 0, Eq. (8-12) now becomes:

/Vl - 0.2242
V -0.224 (0)77
11 = arctan +
(10.9)(57.3) 10.9
arctan(-4.35)
+ 0
624.6

If a calculator is used for determining the Arctan of -4.35, most likely the angle of
-77.05° will be obtained. However, noting that the Arctan expression must lie
between 0 and 180°, we have to subtract -77.05° from 180° to obtain the proper
angle value for the preceding equation. Therefore, we use the angle of 102.95° as
Arctan(—4.35).
Section 8-3 Generating and Analyzing a Transient Response Curve 219

102.95°
U ~ 624.6
= 0.165 sec

To find t2 in Table 8-1, we note that the Arctan expression (i.e., ta in Eq.
8-1 If) doesn’t change. However, we have to calculate a new tb using Eq. (8-1 lg).
We must multiply that equation by an integer, as seen in the second term of Eq.
(8-12), the integer being 1 in this case. Therefore,

(l)?r
t2 — 0.165 +
10.9
- 0.453 sec

The mid-value time, which would correspond to the occurrence of the first peak of
the transient curve located mid-way between the times for n = 0 and n = 1, would
be 0.309.
Our next task is to determine the amplitudes occurring at the three times just
calculated. We do this by using Eq. (8-11). We see that the first and second
crossover times will produce an amplitude of 1 by the definition and intent of Eqs.
(8-11) and (8-12). However, we need to calculate the amplitude of the mid-value
term, t2. Therefore,
^-(0.224X11.2X0.301)
h = 1 + ^.~~^2 sin[(10.9)(57.3)(0.309) - 102.95]

- 1 + sin( 192.99° - 102.95°)

= 1 + 0.473 sin90°
= 1.473
Table 8-1 lists the other crossover time points along with mid-value points
and their amplitudes.
The data listed in Table 8-1 is plotted in Figure 8-9. Notice how long it takes
the pencil in our recorder to finally settle down to read one volt. According to our
transient curve, it takes approximately two seconds or so before it becomes
difficult to see any more maximums or minimums occurring in the curve.

8-3.3 Settling Time

What is actually done in analyzing transient curves is to select a band or range on


either side of the desired set point (one volt in our case) and note the length of time
that it takes the transient curve to come within this band limit. Typically, a band
of ±5% within the set point is selected. In our case, we draw a line at 1.05 volts
and 0.95 volts to form our settling band. Doing this, we see our settling time
becoming approximately 1.2 seconds to within ±5% of the set point. This is just
220 Chapter 8 Transient Analysis
Amplitude (Volts)

an estimate, however. We can actually calculate the settling time using the fol¬
lowing reasoning: Looking back to Figure 2-32 in Chapter 2, we see that after
about three time constants, we are within 5% of being stabilized. We know that
the rate of decay of our curve in Figure 2-32 (curve 2, which coincides with the
behavior of our settling transient curve, as we discover later in chapter 8) is
described by the term e~Z(0nt. Since we agreed that t — 3, then
£> ZCOnt —
e -3

our transient curve from Figure 2-32


Taking the natural logs of both sides of the equation,

Z(nnt = 3

Solving for t, which is now our settling time, we get


3
settling time to within ±5% of set point =- (8-13)
za)n

8-3.4 Number of Oscillations Before Settling

We can also calculate the number of oscillations occurring before completing the
±5% settling time. Since Eq. (8-13) allows us to calculate the settling time to
within ±5% of the steady state time, and since we can find the period of one
damped oscillation of our transient waveform, then settling time -j- period will give
us the number of cycles that can “fit into” this settling time span. In other words,

since 5% settling time =


zcon
Section 8-3 Generating and Analyzing a Transient Response Curve 221

1
and the period t =
r

u>d
and since (od = 2trf, or /
2tt
2tt
then the period for (Od
(Od

We then divide the settling time by the period of our damped waveform to get:
3
Z(on _ laid
2lT Z(On 2TT

U>d
1.5cOd _ l.5o)nVl — z2
Z(On7T Z(On7T

number of oscillations to _ 1.5Vl - z2


(8-14)
±5% of complete settling ztt

For our recorder, the settling time, according to Eq. (8-13), is

3
1.20 sec
(0.224X11.2)

This coincides with the results from Figure 8-8. The number of oscillations before
settling, according to Eq. (8-14), is

1.5 Vl - 0.2242
(0.224)(3.1416)

8-3.5 Over-shoot

Another useful equation allows us to calculate the over-shoot amount. The over¬
shoot of a transient waveform is defined as the ratio, usually expressed as a
percentage, between the amount the first peak of a transient waveform exceeds its
final steady state or set point value. Refer to Figure 8-10. Curve 1 in Figure 8-10
has an over-shoot ratio of 1.5/1 where the 1.5 is the amplitude of the over-shoot
and the 1 is the final steady state value. Expressed as a percentage, the over¬
shoot amount would be 50%. In other words, the curve’s peak amount exceeded
the steady state value of 1 by 50%. Similarly, curve 2 has an over-shoot of 1.2/1 or
20%; curve 3 has no over-shoot, since its ratio of 1/1 gives us a percentage value of
0. To calculate the amount of over-shoot, expressed as a percentage, the follow¬
ing equation may be used (notice that only the curve’s damping factor, z, is
needed to determine the over-shoot):

% over-shoot = e znl^ ^ x 100 (8-15)


222 Chapter 8 Transient Analysis

Amplitude

Response Time —

Figure 8-10 Determining over-shoot and response time.

According to Eq. (8-15), the percent over-shoot amount for our recorder is
(0.224X3.1416)/Vl - 0.2242 x JQQ _ e-0.704/0.975 x jqq

= 48.6%

Note: This result is somewhat larger in value than the result obtained in Table
8-1. The difference in values can be attributed to rounding-off errors and the
method used in calculating the mid-value time of 0.309 sec and its associated
amplitude of 1.473. Theoretically, this value is also the over-shoot value for this
curve according to the calculation methods used.
This is equivalent to an over-shoot amplitude of 1.486 volts on our voltage
recorder. Again, our transient curve plot confirms this amount.

8-3.6 Response Time

The time it takes for a system to go from 10% of its steady state value to 90% of its
steady state value is called the response time. This term can be easily confused
Section 8-3 Generating and Analyzing a Transient Response Curve 223

with a system’s time constant. These are two entirely different terms. Unfortu¬
nately, many reference books and textbooks confuse these two terms by using
them interchangeably. If you refer again to Figure 8-10, you see the response time
term being graphically explained. Response time is used extensively by servo
manufacturers. The reason for omitting the first 10% and last 10% of a perfor¬
mance curve is to avoid any initial or final disruptions or deviations in an other¬
wise linear or smooth curve, which occasionally happens in the graphing of the
performance characteristics of many devices.
Now that we have defined the many important features of a transient func¬
tion, we can relate some of this information to our earlier closed-loop frequency
response curves. Referring back to Figure 8-6, we note that the frequency of the
peak of our gain curve occurs at approximately 11 rad/sec, or 1.75 Hz. Knowing
now the z and con of our system, we can calculate this closed-loop peak fre¬
quency, com.

ojm = ojn s/ 1 2z2 (8-16)


In our case, com = 11.2V 1 - (2)(0.2242) = 10.62 Hz. Knowing this calculated
value, we can now insert it into Eq. (8-4) and calculate the closed-loop peak gain,
Am, and the resultant phase angle.

A
25 25
25 + ./10.62 - (0.2)(10.62)2 ' 25 - 22.56 + 7*10.62
25 25
= 2.29 A—77.06°
2.44 + J10.62 “ 10.9 Z77.06

Converting the gain of 2.29 to decibels, 201og2.29 = 7.2 dBA. Therefore, the
maximum peak is 7.2 dBA at a phase angle of -77.06°.
We can also calculate the bandwidth, cob, of our system instead of relying on
our response curves for this information:

(0b = 0>„Vl - 2Z2 + V2 - 4z2 + 4z4 (8-17)


In our case,

cob = II.2V1 - 2(0.224)2 + V2 - 4(0.224)2 + 4(0.224)4


= 16.78 rad/sec

We now have a design problem that we must resolve with our chart recorder
system. How do we correct the jitteriness of our marking pencil? Looking at the
open-loop Bode plot for our system (Figure 8-5), we see that we need to increase
our phase margin somewhat in order to deaden the recorder’s responsiveness.
One obvious approach is to reduce the gain of our system. This can be done by
either reducing the amplifier gains or by increasing the gear ratio on the transmis¬
sion box attached to the servomotor, or a combination of both. Or, we could
increase the viscous damping on our servomotor in an attempt to dampen the
oscillations. Whatever we decide to do, though, we must keep in mind that our
224 Chapter 8 Transient Analysis

system’s bandwidth will be affected. If we want a system that is capable of


responding to rapid fluctuations in voltage input, then the bandwidth should be
kept as high as possible so the servomotor can respond accordingly. If we aren’t
interested in catching quick voltage changes, then we would probably be satisfied
with the present bandwidth, or something even smaller.
To approach this problem, let’s first look at Eq. (8-17) again. Notice that if
you make the value of z grow larger in an effort to deaden a system’s response, the
value of ajb becomes smaller. Our Bode plot bears this out. As you attempt to
reduce the gain in an effort to increase the phase margin, the open-loop bandwidth
decreases. Notice what happens as the gain curve in Figure 8-5 is lowered. The
zero gain crossover point shifts to the left causing the phase margin to increase,
but at the same time, causing the bandwidth to decrease. It’s reasonable to as¬
sume that if the open-loop bandwidth decreases, so will the closed-loop band¬
width. So it seems that the only really logical thing to do, without going through a
major redesign effort on our hardware, is to try to decrease the system’s gain and
hope that this will not affect our bandwidth all that much.
Looking at Figure 8-5, if we were to lower our gain curve by, say, 16 dBA,
this would give us a phase margin of 49°. Notice too, that as the gain curve is
lowered by this amount, this shifts us from the -40 dBA/decade slope of the
curve and puts us on the -20 dBA/decade slope instead. This is a good place to
be. A rule of thumb in controls design is, operate on the -20 dBA/decade slope of
a gain curve whenever possible. This insures good system stability.
A decrease of 16 dBA is the equivalent of reducing our system gain by a
factor of 0.158 (201og. 158 = -16dB). The easiest way to do this is merely to
reduce the gain on one of our two amplifiers on our recorder. The resulting open-
loop transfer function now becomes, 3.96/5(1 + 0.25). See Figure 8-11. The
closed-loop transfer function becomes

3.96 3.96
KG _ 5(1 + 0.25) _ 5(1 + 0.25)
1 + KG ~ 3.96 ~~ 5(1 + 0.25) + 3.96
+ 5(1 + 0.25) 5(1 + 0.25)
3.96 _ 3.96
5(1 + 0.25) + 3.96 “ 5 + 0.252 + 3.96
3.96 3.96
(8-18)
~ jto + 0.20'w)2 + 3.96 ~ 3.96 + jco - 0.2a2

Our next task is to determine the value of z. This will tell us for sure whether or
not our system is going to be stable. Also, from z, we can calculate what our
bandwidth is going to be without having to plot a closed-loop frequency response
curve. Following the procedure that we performed earlier, we use Eq. (8-5) as a
pattern to help us determine our new value of z. Using Eq. (8-18),
Section 8-3 Generating and Analyzing a Transient Response Curve 225

PHASE ANGLE (DEG)


Figure 8-11 Bode plot for expression 3.96/5(1 + 0.25).

1 „ 3.96
5 X 0.2s2 + s + 3.96

1
0.2s2 s
+ + 1
3.96 3.96

x
19.61
0.051s2 + + 1
3.96 19.61

s" 19.61s
+ + 1
19.61 77.66

1 1
+ (19.61)(0.057)s + s2 1.118s
s
19.61 4.43 19.61 + 4.43 +
> (77.66X0.057) = VW6\ = 4.43 (8-19)
226 Chapter 8 Transient Analysis

Since 2z = 1.118 (in Eq. 8-19), z = 0.559. This value of z indicates that we now
have a stable system. Our new con is 4.43 rad/sec. Now that we know the z and
(on, we can calculate the new bandwidth using Eq. (8-17):

03fj = 4.43 Vl - 2(0.559)2 + V2 - 4(0.559)2 + 4(0.559)4


= 5.32 rad/sec

This is a substantial reduction in bandwidth compared to the earlier bandwidth


result when we had the more jittery system. Converting the 5.32 rad/sec to its
frequency equivalent (using Eq. 1-22 from Chapter 1) we obtain an equivalent
frequency of 0.85 Hz. This means that our recorder will not be able to record
voltage events at their proper magnitudes that change any more rapidly than about
0.85 Hz. We can now hand-plot the transient response curve using Eqs. (8-11)
and (8-12). The resulting curve along with the calculated crossover times are
shown in Figure 8-12. We can check our work by calculating the ±5% settling
time and the oscillation numbers to settling:

3
±5% settling time = — (from Eq. 8-13)
zcon

3
~ (0.559)(4.37)
= 1.228 sec
Amplitude (Volts)

Figure 8-12 Transient response curve for recorder whose transfer function is
3.96/0.2 s2 + j + 3.96.
Section 8-4 An Example Problem 227

To determine the number of oscillations:

number of oscillations = —-— (from Eq. 8-14)


Z7T
_ 1.5Vl - 0.5592
(0.559)(3.1416)
= 0.708

These calculations agree with the data plotted in Figure 8-12.


As a result of decreasing our gain, we gained stability in our system but we
sacrificed speed of response. It now takes over one second for our recorder to
reach its final set point goal of one volt.

8-3.7 Handling Other Values of Step Inputs

Up to this point, we have dealt with a system response that has been given a step
input quantity of 1. In the case of our voltage recorder, we subjected the recorder
to an input of one volt to see how it would behave. But we have to ask ourselves,
what about other step input values other than 1? Are they treated any differ¬
ently? The answer is no. They are treated just like the unit step input. The only
thing that’s different are the amplitudes involved in generating the transient re¬
sponse curves. As an example, assume that we placed a 3.5-volt DC signal across
our voltage recorder’s input. Assume also that we are using the same system
configuration that we just finished analyzing. How will our transient curve differ
from the one in Figure 8-12? To answer this question, let’s take another look at
Eq. (8-10). This equation is set up for a unit step input quantity. The 1 in the front
of this equation causes our plot to be generated around a horizontal axis of 1. We
find out later in this chapter that the exponential expression, e~za)nt + 1 is the
decaying “envelope” of our damped oscillating wave, which is our transient
waveform (see Figure 8-13). We can change the magnitude of this function and
shift the horizontal axis by merely multiplying the entire function of Eq. (8-10) by
a constant, A. Therefore, A in our case will be the 3.5-volt step input. Our resul¬
tant waveform, as a result, will increase in magnitude by 3.5 times and become
generated around a horizontal axis of 3.5 instead of the usual 1 axis. None of the
times or frequencies that we calculated earlier will change. This is similar to the
case of working with ordinary sinusoidal waveforms. Changing the amplitude of a
sine wave doesn’t affect its period or frequency.

8-4 AN EXAMPLE PROBLEM

Now that we have discussed most of the important features of a control system’s
transient behavior and can calculate many of the more useful parameters, let’s go
228 Chapter 8 Transient Analysis

(a)
Amplitude

Figure 8-13 The components of the


transient waveform:

(w„Vl - z2t - <M, 1 e~Z(,,nl sin


V1 — zl

where cr = a>nt, K — —.
vr=7, and <t
Vl — z2
= arctan -
—z

through some additional sample calculations. We use one of our closed-loop


transfer function examples that we discussed back in Chapter 7.

EXAMPLE 8-1
Using Eq. (7-1) and assuming this is a closed-loop transfer function describing an
automatic positioning system, calculate the following transient function features for
this equation. Assume a step input of 2.7 volts. Find:
Section 8-4 An Example Problem 229

a. damping factor
b. the natural frequency

c. the damped frequency JV20 25 + > + 100 (Eq‘ 7'^


d. ±5% settling time
e. maximum over-shoot
f. oscillation number to ±5% settling
g. the bandwidth
h. the closed-loop peak frequency
i. the closed-loop peak gain and phase angle

Solution:
a. To find damping factor (z):
Convert to a step input Laplace transform using Eq. (8-5):
1 1
(Eq. 8-5) =
s* . 2zs 0.25s2 5
+ + 1
-CO* CO, 100 + 100 +

1
(8-20)
^2 0.25
15,2 5 1 5
Uoo + 100 + Uoo + 20 + *J
From Eq. (8-20), we see that 2z = 0.2; therefore, z - 0.1. Also,
b. To find (on, we again look at Eq. (8-20), which says that con 20 rad/sec. -

c. To find (od, we use Eq. (8-9):


ojj — wnVl - z2 = 20 Vl - 0.12 - 19.9 rad/sec
d. To find ±5% settling time, we use Eq. (8-13):
3 3
:5% settling time = = 1.5 sec
zcon (0.1)(20)
e. To find the maximum over-shoot, we use Eq. (8-15). However, since this is a
percentage over-shoot value that will be given us, we have to multiply this
percentage by the step input amount to find the actual amplitude of this first
over-shoot oscillation. Therefore,
over-shoot = e -zirly/l
v‘ - z2 — g-(0.1)(3.1416)/Vl - 0.12
* = e
= 0.729
The amplitude of the over-shoot = 2.7 x 0.729 = 1.97 volts,
f. To calculate the number of oscillations, we use Eq. (8-14):

1.5Vl - z2 1.5 Vl - 0.12


number of oscillations =
Z7T ~ (0.1X3.1416)
= 4.75
g. The closed-loop bandwidth is found by using Eq. (8-17):
230 Chapter 8 Transient Analysis

cob = V'l - 2z2 + V2 - 4z2 + 4?_


= Vl - 2(0.12) + V"2 - 4(0.12) + 4(0.P)
= V2J8
= 1.54 rad/sec

h. The closed-loop peak frequency is found by using Eq. (8-16):

com = conV 1 - 2z2 = 20 Vl - 2(0.12) = 19.8 rad/sec

i. To find the closed-loop peak gain, Am, and its associated phase angle, we plug
the value of our just calculated a)m into the closed-loop transfer function, Eq.
(7-1).

100
A —
-(19.8)2(0.25) + ./(19.8) + 100
100
= 5.03 A-84.3°
-98.01 + 100 + J19.8
- 14.03 dBA A-84.3°

Based on all the calculations that were just completed, you can see that in
many instances it wouldn’t be necessary to plot the response curves for a control
system. Much of the needed information for evaluating a system can be obtained
through these kinds of calculations instead.

8-5 THE COMPONENTS OF A TRANSIENT WAVEFORM

If we look closely at Eq. (8-10), we can see that this equation is the product of two
components. We see an exponential expression, Ke~at, (where K and cr are con¬
stants) and a sine expression, sin(co^ - <£>) (which looks very much like the
expressions discussed in Section 1-7.3 of Chapter 1). Figure 8-13(a) shows the
individual components; Figure 8-13(b) shows the combined results. Based on this
approach, we could rewrite Eq. (8-10) into the following form:

y = Ke~atsm((e>dt - <j>) (8-21)

It’s easy to see that the Ke~at expression acts as an envelope that encapsulates the
sinusoidal expression. This is the decaying function portion of the transient wave¬
form curve that is often mentioned in transient analysis discussions. Figure 8-14
shows the relationships between the various equations used to calculate the differ¬
ent quantities of the transient waveform.

8-6 THE ROOT LOCUS METHOD

Another method that is frequently used among control system engineers for the
analyzing of transfer functions is much more mathematical than the previous
Section 8-6 The Root Locus Method 231

a ~ z2
0 = arctan -
-z

Q-z-n/yJ] - z2

Figure 8-14 The components of a transient waveform.

method and requires some preliminary ground work in dealing with exponential
expressions and trigonometric functions. This mathematical approach in analyz¬
ing transient behavior is called the root locus method.

8-6.1 Euler’s Equation and the Complex S-Plane

To help us understand the root locus method, we have to go back to the two
transient components of Eq. (8-21). We see a sine function mixed with an expo¬
nential function creating a rather confusing picture of the type of expression we
are dealing with. It would help us if we could convert one or the other expression
into the other so that we would have either all exponential terms or all trigonomet¬
ric terms.
There is an identity that allows us to convert cosine expressions to powers
of e. It is:
eid = cos 0 + j sin# (8-22)

This-equation is called Euler s equation, named after the Swiss mathematician,


Leonhard Euler (pronounced Layonhart Oiyler). The proof for this expression is
somewhat extensive. However, a proof is presented in Appendix D.
Now, let’s look at another form of Euler’s equation. We can also state that

e~je = cos0 - jsind (8-23)


232 Chapter 8 Transient Analysis

If we add Eq. (8-22) to Eq. (8-23), we will have


eje _|_ e-je = 2cos 6

pj® -i-

or -2-= cose (8-24)

And since 6 = cot, then


-j- p
---= cos a)t (8-25)

We now have an expression that allows us to change any cosine expression


into an exponential expression. Notice that it takes both +ja>t and -ja>t to make
one cosine curve.
What we have just proven is fine for converting cosine expressions into
ex-typc expressions, but Eq. (8-21) has a sine expression in it, not a cosine expres¬
sion. However, since the two expressions are related by a phase shift of 90°, we
can, for all practical purposes, say they are equivalent as long as we remember to
make the 90° phase adjustment. Furthermore, in Eq. (8-21), in place of the
codt — </> expression, let’s substitute the general expression 6. What we have now
is a very generalized transient expression, Ke~atcosS (also, let’s assume that
contained within our 0 expression is our 90° sine-to-cosine phase shift conver¬
sion). Now, we can make a further substitution using Euler’s equation so that we
have

(ejd + e~je)
Ke (8-26)
2
Let’s leave our discussion on Euler’s equation momentarily and focus our
attention on this next concept: Notice the similarity in appearance between Eq.
(8-22) and the complex expression r = a + jb. Equation (8-22) appears to be a
complex expression describing the location of a point, eje, drawn on a set of real
and imaginary axes. This is depicted in Figure 8-15. The similarity between this

Imaginary Imaginary

(a 4- jb)

Real

Figure 8-15 The similarities between a + jb and cos# + jsmO.


Section 8-6 The Root Locus Method 233

and the complex number, r = a + jb, should now be even more apparent. Let’s
construct a set of real and imaginary axes and label the imaginary axes as 5 (or jco)
and the real axes as cr (or cont, since <x = cont). If we were to let <x = 0 and assign a
wide range of values to 6 (cot), we would create a whole series of sinusoidal
waveforms whose frequencies would vary depending on the values of 6. These
curves, if superimposed on our s-plane axes, would appear along the vertical axis
of our 5-plane plot as shown in Figure 8-16.

/'<*>

Figure 8-16 How cosine waves and


their frequencies would position them¬
selves along the ja> or s axis of the s-
plane. (Originally presented in B.E.
DeRoy’s Automatic Control Theory,
John Wiley and Sons, publisher, 1966,
and reproduced here with the pub¬
lisher’s permission.)

Looking once again at Eq. (8-26), let’s now see what happens when we let
6 = 0 and allow or (cont) to vary over a wide range of values. Again, we superim¬
pose the results onto our 5-plane. Figure 8-17 shows us the outcome. Notice how
the exponential slopes (representing the envelopes of our transient waveforms)
increase and decrease the further you move away from the right or left of the
5-axis. What we are interested in are the combined results of Figures 8-16 and
8-17 since, after all, the expressions depicted in these two figures are in fact
multiplied together according to the original math expression of Eq. (8-26). Figure
8-18 shows us the final results. Notice how, as you move to the right of the jco- or
5-axis, the resultant transient waveforms become increasingly unstable. It’s only
to the left of the 5-axis where stability seems to exist.

8-6.2 The Characteristic Equation

Now that we have developed the concept of the complex 5-plane and how it
can help us to visualize the stability characteristics of our transient waveform, we
put this concept to practical use.
Assume we have an open-loop transfer function that looks like the following:

flout _ (s + 3)
(8-27)
Fjn (S + 1)(5 + 4)
234 Chapter 8 Transient Analysis

/<*j

Figure 8-17 The results of letting 6 =


0 and allowing a to vary, resulting in
curves seen along the cr-axis. (Origi¬
nally presented in B.E. DeRoy’s Auto¬
matic Control Theory, John Wiley and
Sons, publisher, 1966, and reproduced
here with the publisher’s permission.)

The denominator of this expression, (5 + l)(s + 4), is called the characteristic


equation for Eq. (8-27). The reason for this is because the denominator contains
all the important behavioral characteristics for the entire equation. We can prove
this by the following method. Temporarily ignoring the numerator of our open-
loop transfer function, we can break up the transfer function into two separate
fractions, each having its own numerator of, say, A and B. Then we can recom¬
bine these two fractions back into a single fraction as shown:

(■s + 3) = A B = A(s + 4) + B(s + 1) ,


(s + l)Cs + 4) (s + 1) (s + 4) (s + l)Cs + 4) 1 }

/CO

A/V AAA

Figure 8-18 Multiplying together the


results of Figure 8-16 and 8-17. (Origi¬
nally presented in B.E. DeRoy’s Auto¬

y/V' aAAj rx^\j matic Control Theory, John Wiley and


Sons, publisher, 1966, and reproduced
s Plane here with the publisher's permission.)
Section 8-6 The Root Locus Method 235

We now equate the numerator in Eq. (8-27) to the resulting numerator in Eq.
(8-28):

5 + 3 = A{s + 4) + B(s + 1) (8-29)


If we let 5 = — 1 in Eq. (8-29), then

-1 + 3 = A( — 1 + 4) + B( 0)
and 3A = 2
or A = §

If we let s = 4 in Eq. (8-29), then

-4 + 3 = A(—4 + 4) + B{-4 + 1)

3B = 1
and B =

Placing the values of A and B back into Eq. (8-28), we get:

(5 + 3) _ A(5 + 4) + B(s + 1)
(5 + 1)(5 + 4) (5 + l)(s + 4)

\ {S + 4) + \ + 1) (f) 5 + I + (1) 5 + l
(5 + 1)(5 + 4) (5 + 1)(5 + 4)
5 + 3
(5 + 1)(5 + 4)

which is what we started with.


While this is certainly not a rigorous proof that all denominators control the
behavior of the entire equation, we can certainly see that this is the case for the
example just given. However, this is true for all transfer function equations. In
algebra and analytic geometry, the numbers -1 and -4 are called the characteris¬
tic equation’s roots. Any numeric values that cause a fractional equation’s de¬
nominator to go to zero (or, to state it another way, any numeric values that cause
the entire equation to go to infinity), are called roots. The roots are referred to as
being poles for the entire equation. On the other hand, any numeric values that
cause the numerator to go to zero, and consequently, the whole fractional value to
go to zero, are called zeros. In the preceding case, -3 would be that equation’s
zero.

EXAMPLE 8-2 -
Find the poles and zeros for the following expression:
6out 23.8(1 + 0.75)
Em ~ 5(1 + 0.45)0 + 0.55)
236 Chapter 8 Transient Analysis

Solution:
Finding the roots of the characteristic equation would result in identifying the poles
for the entire equation. Consequently, solving for 5 in the first parenthesis of the
characteristic equation: 0.4s = -1, therefore, 5 = -2.5. Solving for s in the second
parenthesis: 0.5s = — 1, therefore, s = -2. And don’t forget the s. Letting s = 0 will
also cause the denominator to go to zero. Therefore, s is also a pole. As a result, the
poles for the given transfer function are -2.5, -2, and 0.
To find the zeros, we solve for s in the parenthesis in the numerator. There¬
fore, since 0.7s = — 1, then s = -1.429. The zero for the given transfer function is
-1.429.

The reason why we want to know the poles, or roots, of a transfer function is
this: Any positive roots occuring in that equation are a definite indication that that
particular system is absolutely unstable. Under no circumstances can stability
exist. However, we cannot say that because there are no positive roots present in
our equation, the system is absolutely stable. On the contrary, the system could
still be unstable, or jittery, because of too low of a z value. In this case, it would
be mathematically proper to say that because no positive roots are present, the
system is possibly stable.
Just one positive pole is all that it takes to make a system absolutely unsta¬
ble. Recalling our discussion earlier concerning the complex 5-plane, we con¬
cluded from Figure 8-18 that any plotted (cr, 5) points lying to the right of the 5-axis
will result in an unstable system. A positive pole is one such point. While this is a
simple enough observation to make, it’s the determining of the poles that can
sometimes be difficult. This is the case when the expression is not in its factored
parenthesis form but is, instead, in the unfactored or expanded polynomial form.
(A polynomial form is any algebraic expression that is a sum of terms of the form,
axn, where the n in each term is a positive number, or power, usually arranged in
descending order from one term to the next.) For instance, if Example 8-2 were
stated in the form,

$out _ 23.8(5 + 0.7)


Em 0.253 + 0.952 + 5

instead of the given factored form, you would probably have more difficulty in
determining the roots. You would first have to factor out the 5 term in the denomi¬
nator and then, either by inspection or by using the quadratic equation, determine
the other root values.
Pole determination becomes increasingly difficult as the order of the charac¬
teristic equation increases. Look at the following equation containing a third-
order expression:

0out _ 2170(1 + 0.255)


Em ~ 353 + IO52 - I65 - 32
Section 8-6 The Root Locus Method 237

Factoring this expression into its roots could be a time-consuming chore. Fortu¬
nately, there is an easier way.

8-6.3 The Routh-Hurwitz Stability Test

There are two cases of unfactored polynomial characteristic equations that we can
automatically assume to have positive roots. As a result, these cases generate
points in the right-hand portion of our 5-plane, resulting in absolute instability.
The first case has to do with those polynomials containing minus signs, such as
5v3 + 4x2 - 9, or \6x4 - 32. The second case has to do with those polynomials
having missing terms, such as x4 + 3x3 + 7, or 6v3 + 5. Factoring, as in the first
case, will produce either positive real-number roots or roots containing complex
numbers with positive real numbers. Either case will produce points in the right-
hand portion of the 5-plane and are, as a result, absolutely unstable.
If determining stability is the only thing we are interested in, and we are not
interested in the root values themselves, then all we need is some method that
would allow us to determine only the signs of our roots. The Routh-Hurwitz
Stability Test (pronounced Rooth-Herwitz) is one such method. To use this
method, we first note that our polynomial is of the general form.
a0sn + a\Sn~x + ci2Sn~2 + a^sn~3 + • • • + an = 0 (8-30)

We then arrange the coefficients associated with each term into the following
pattern or matrix:
ao a2 a4
ax a3 a5
b\ b2 b2
Cl c2
dx (8-31)

The terms beyond a5 must be calculated using the following equations:


CliClj ci\d4 — a0a5 Cl\Cl() <3o<27
h\ - ^3 ai
a\ ai
b\ai ~ ci\b2 b\as ~ a\b2 c\b2 ~ bxC2
Cl = c2 = d\ (8-32)
bx Cl

Our purpose is to fill out the terms in the far left-hand column of the array. If
these terms are all positive and nonzero, there are no positive roots existing in the
original equation and the system is possibly stable (i.e., it’s not absolutely unsta¬
ble). Let’s try a couple of examples.

EXAMPLE 8-3 -
Determine the stability characteristics for the function:

25
0.2F + 5 + 25
Solution:
This is the closed-loop transfer function whose transient waveform was plotted in
Figure 8-9. The computed damping factor was 0.2; the system was not stable, but
jittery. Using the Routh-Hurwitz Test: a0 = 0.2, a\ = 1, and a2 = 25.
Calculating b\:

, (1)(25) - (0.2)(0)
b' = -T-
- 25

All the other factors contain zeros; consequently, all factors beyond b\ do not exist.
Finally, we arrange our factors according to the array, Eq. (8-31):

0.2 25
1
25

Inspecting the left-hand column, we see that there are no negative terms; therefore,
there are no roots in the right-hand 5-plane and our system is possibly stable. (Of
course, we knew that it was somewhat unstable to begin with, but the Routh-
Hurwitz Stability Test verifies this by implying that this equation is possibly stable.)

EXAMPLE 8-4 “-
Determine the stability characteristics for 53 + 52 + 2s + 8.

Solution:
Setting up the Routh-Hurwitz array, we get: a0 = 1, a\ = 1, a2 = 2, a3 = 8, and
a4 = 0. Any a-terms needed beyond a4 would also have 0 values. Calculating the
other factors, we get:

, (1)(0) - (1)(0)

(1)(2) - (1)(8) (-6X8) - (1)(0)


ci =

= -6 = 8

Our array now looks like:

1 2
1 8
-6
8

Because of the one negative term, -6, our equation is absolutely unstable. Note:
You can determine the number of positive roots in the original equation by merely
counting the number of sign changes that take place within the left-hand column. In
the preceding case, there are two sign changes; one change is going from + to —, and
the other from - back to +. Therefore, there are two positive roots. We can con¬
firm this by noting the actual roots of our equation. They are 5 = -2, 5 = 1/2(1 +
j3.87), and 5 = 1/2(1 — j3.87). The complex roots are both positive, since the real
number portion (the 1 in both cases) is positive.

238
Section 8-6 The Root Locus Method 239

8-6.4 The Root Locus

The preceding discussion brings us to the concept of the root locus. When we talk
about a locus of a moving object, we are speaking about a path of that object. For
example, if you were to install a small light onto the rim of a wheel and then roll
the wheel along a flat surface, the locus that the light would make on a time
exposure photograph would be a series of looped spirals. This is the concept that
we analyze here. We look at a time exposure showing the path that a series of
plotted points will make on a graph as certain variables are changed within an
equation.
In this section we investigate the path or locus that the roots of a given
transfer function make on the 5-plane as a quantity within the equation is
changed. That quantity will be the gain, contained within the constant usually
associated with a transfer function’s numerator.

First-Order Functions:

Let’s take a look at the expression:

70(1 + 0.25)
KG (8-33)
1 + 0.55

Equation (8-33) is an open-loop transfer function for a phase-lead network being


used in conjunction with an amplifier. The 70 term is comprised of the amplifier’s
gain and also the resistor ratio, Ri!{R\ + Ri), normally associated with a phase-
lead network’s transfer function. Equation (8-33) has a pole, 5 = — 2 and a zero,
5 = —5. Converting Eq. (8-33) to its closed-loop counterpart, we get:

KG _ 70(1 + 0.25)
(8-34)
1 + KG 1 + 0.55 + 70(1 + 0.25)

Looking at the characteristic equation, we solve for 5 to determine its roots. We


get 5 = -4.90. Figure 8-19 shows the plots for both the poles and zeros (shown as
an x and o respectively) from the open-loop expression of Eq. (8-33) and the one
root (shown as a •) from the closed-loop expression, Eq. (8-34).
We now want to observe what happens when we vary the gain of our system
in Eq. (8-34). In other words, instead of having a fixed value of 70 for our gain,
let’s make that our variable now and assign various values to it. We can then
observe what happens to the root in the characteristic equation that we have
plotted on the 5-plane. Equation (8-34) now becomes:

KG _ Kx{ 1 + 0.25)
(8-35)
1 + KG 1 + 0.55 + Kx( 1 + 0.25)

Looking at the characteristic equation and solving for 5, we get:

-Kx - 1
s (8-36)
0.5 + 0.2 Kx
240 Chapter 8 Transient Analysis

Figure 8-19 The poles, zeros, and roots for 70(1 + 0.2s)/l + 0.5s and 70(1 +
0.2s)/l + 0.5s + 70(1 + 0.2s).

We now pick a range of values, beginning with zero for Kx, and record the
resultant values of 5. Table 8-2 shows our results.

TABLE 8-2 VALUES OF THE


ROOT 5, RESULTING FROM
ASSIGNING VARYING VALUES
OF Kx TO THE CHARACTERISTIC
EQUATION, 1 + 0.5s + Kx( 1 + 0.2s)

Kx s

0 -2
1 -2.22
5 -4
10 -4.4
20 -4.67
50 -4.86
100 -4.93
Section 8-6 The Root Locus Method 241

Figure 8-20 shows the results of plotting the roots from Table 8-2. Notice
how, beginning with a Kx of 0, the path or locus of our roots begins at —2, the zero
of our open-loop function, and moves toward the final value of -5, the pole. The
arrows indicate the path movement. This is typical of any first-order characteris¬
tic equation. Notice too, that there are no points located in the right-hand plane
of our plot. This is the absolutely unstable region of the plane. This tells us that
all first-order equations are possibly stable. There are no absolutely unstable
systems when dealing with first-order systems.

Figure 8-20 The root locus of a first order equation showing migration of roots
from zero to pole.

Second-Order Functions:

Let?s now analyze a second-order system. Looking back at Section 8-3 where we
experimented with our voltage chart recorder, we varied the gain of our control
system in an effort to stabilize it. We lowered the gain by reducing the value of
the constant, 25, in the closed-loop transfer frunction, 25/(0.2s2 + 5 + 25). This
obviously had an effect on the roots of the characteristic equation, 0.2s2 + s +
242 Chapter 8 Transient Analysis

25. The reason for this is due to the fact that as the 25 is altered in the open-loop
function, 25/5(1 + 0.2s), the characteristic equation in the closed-loop function is
also varied. That same constant, whatever value it may be, always shows up as
the coefficient for the quadratic equation (the second-order expression) in the
denominator that results from performing the open-loop-to-closed-loop con¬
version.
Let’s determine the roots of 0.2s2 + 5 + 25. Using the quadratic formula,
5 = [ — b ± Vb2 — 4ac]/2a, we find that the roots are S\ = —2.5 + jl0.9 and
52 = —2.5 — 710.9. Graphically, we can represent these roots as points on the <j

versus 5 coordinate 5-plane axes as seen in Figure 8-21. (Locating these points on

s(jco)

• 11
(-2.5 + y 10.9)
10

9 —

8 —

7 —

6 —

5 —

3 —

1
1 1 1 1 1 1 I I i I i i
CM

T
CD

CO
ID

1
1

1
1

-2

-3

-4

-5

-6

-7 -

-8

-9

-10
(-2.5 - / 10.9)
• -11 —

Figure 8-21 Root location for s = — 1 + j4.36 and s = — 1 — y‘4.36.


Section 8-6 The Root Locus Method 243

the coordinate 5-plane as we just did is similar to locating impedances on the R,j
axes of a phasor diagram. If necessary, see Appendix A to review how this is
done.) As we have done before according to mathematical custom, the real num¬
bers in our complex expressions are located along the horizontal axis; the imagi¬
nary j expressions are plotted along the vertical axis.
Now that we have located the roots of our characteristic equation, let’s look
at a more generalized expression that states:

Kx
KG = (8-37)
5(1 + 0.25)

In place of the constant, 25, we have substituted Kx. This will result in a closed-
loop characteristic equation of 0.252 + s + Kx. Next, let’s plug in a range of
values for Kx, just as we did for the first-order example, and for each one of these
values that we choose, calculate the roots of that particular characteristic equa¬
tion. Table 8-3 lists our results.

TABLE 8-3 DETERMINING THE


ROOT LOCUS FOR 0.2s2 + s + Kx

Kx s \ root s2 root

0 -5 0
1 1.38 -3.62
2 -2.5 + 71.94 -2.5 - 71.94
3.96 -2.5 + 73.68 -2.5 - 73.68
10 -2.5 + 76.61 -2.5 - 76.61
25 -2.5 + 710.9 -2.5 -710.9
50 -2.5 + 715.6 -2.5 - 715.6
100 -2.5 + 722.2 -2.5 -722.2
300 -2.5 + 738.6 -2.5 -738.6
500 -2.5 + 749.9 -2.5 - 749.9
700 -2.5 + 759.1 -2.5 - 759.1
1000 -2.5 + 770.7 -2.5 - 770.7

Now, we’ll plot our results on 5-plane coordinate axes. Figure 8-22 shows
the results. Notice the arrows indicating the path of travel of these roots as the Kx
value increases from 0 to a very large value. These paths are the root loci and are
typical of all root loci for second-order characteristic equations. As we noticed in
our first-order expression, the root locus begins at the pole of our open-loop
expression when Kx = 0. All closed-loop root-locus plots for second-order equa¬
tions begin at the open-loop poles and progress toward their mid-point as Kx is
varied from 0 to infinity. In the case of our closed-loop expression, the mid-point
happens to be the coordinates (-2.5, 0) on the 5-plane. After the loci reach this
mid-point, they split, one traveling vertically upward, the other traveling verti-
244 Chapter 8 Transient Analysis

Figure 8-22 The root locus for the characteristic equation 0.2s2 + s + Kx.

cally downward. It’s important to keep in mind what is happening here: We are
varying the gain of our system’s amplifier to give us the various values to our Kx.
We could have just as properly, but not as easily, varied the gear ratio, Kgear, or
the servomotor’s velocity constant, Kve\. This is because Kx = Kgear x Kve\ x
Kamp. Varying the gain of an amplifier is so much easier to do. This can be done
Section 8-6 The Root Locus Method 245

simply by varying a control knob or a screw-adjusted pot somewhere on the


amplifier’s chassis or printed circuit board.
Notice in Table 8-3 that we included the Kx values of 25 and 3.96. This was
done purposely, since these values were used earlier in our analysis of the voltage
recorder. Let’s take a closer look at these two points, first the Kx value of 25.
This point has been replotted, using the coordinates of (-2.5,7'10.9), on an ex¬
panded 5-plane coordinate axis in Figure 8-23. We will ignore the (-2.5, —710.9)
point, since it is a mirror image of the other point. However, note the numeric

value of the 7-value that we did plot, (the 10.9 value). This happens to coincide
with the (od value that we calculated for our system when we analyzed it earlier in
Section 8-3.1 for a Kx of 25. Now, if we think of our plotted point as an impedance
in an electrical circuit and perform a rectangular-to-polar conversion, just as you
would do in an electrical circuit, the value of the r vector is con, and the resultant
phase angle, 6, is the arccosine of our damping factor. According to the data that
we gathered for this system in Section 8-3.1, con = 11.2 rad/sec and z = 0.224.
From our root locus plot in Figure 8-23, ojn = 11.2 rad/sec and z = cos 77.08° =
0.224.
We can prove to ourselves that z = cos# by the following relationships: If we
let x = the length of the horizontal base of the triangle formed by the other two
sides of cod and <*>„, then by the Pythagorean theorem:

"n = "rf + *2

then' JC
2 —
- Q)
,.,2 _ co:
, .2

or = VoF — oA

If z = cos# =
^- °>J
GJn
246 Chapter 8 Transient Analysis

then, cross-multiplying: V(i)2n - 0J2d - ZOJn

Clearing the radical: u2n- 0>2d = zWn

Solving for a)d: COd = V(iJ2n - z2(02n

resulting in: wd = wnV 1 - Z2

This result checks with Eq. (8-9).


Hopefully by now, you have been able to see the advantage of using the root
locus method. It enables a person to trace out the root locus of his or her particu¬
lar characteristic equation so that the Kx value can be varied to see what value
produces the most stable system. The root locus gives a good visual representa¬
tion of what is happening within the control system.

Third-Order Functions:

Let’s look at the situation involving third-order systems. These open-loop trans¬
fer functions are typically of the form:

fl(3Ut _ KC _ _ KXG ^l)_


(8-38)
Em (1 + 5t2)(1 + 5t3)(1 + st4)

The closed-loop equivalent of this expression would be:

flout = KG = _Kx( 1 + ST i)_


E[n 1 + KG (1 + st 2)( 1 + 5t3)(1 + ST4) + Kx{ 1 + st 1)

The third-order characteristic equation for this expression looks somewhat men¬
acing to handle. But there is at least one shortcut that we can look for. If any one
of the time constant values in the denominator of Eq. (8-39) is reasonably close in
numeric value to the time constant in the numerator, we can cancel out those
close terms and wind up dealing with only a second-order equation. What we
mean by reasonably close is that the zero value in the numerator is no more than
about two times greater than any one of the pole values in the denominator of the
open-loop expression. After cancellation, we are left with a second-order equa¬
tion that can be handled in the usual way as far as determining its root locus. This
will still produce reasonably accurate results. Look at the following example.

EXAMPLE 8-5 --
Find the damping ratio for the following expression using the root locus method:

450(1 + 0.25)
KG
(1 + 105)(1 + 0.255)0 + O.85)

Solution:
The zero for the preceding expression is -5 (0.25 = 1, therefore 5 = -1/0.2, or -5).
Section 8-6 The Root Locus Method 247

The three poles of our expression are —0.1, -4, and —1.25. Since the zero value of
-5 is reasonably close to the one pole of -4, we can cancel these two expressions
containing those values, resulting in a final expresion of:

(1 + 10j)(1 + 0.85)

We can now proceed with plotting the root locus of our finalized expression treating
it like any other second-power system expression. In reality, the results will be quite
close to the results obtained if we had actually plotted the root locus for the original
third-order equation. We won’t complete the solution for the damping ratio here; we
only want to demonstrate the canceling procedure just discussed. Otherwise, the
solution would proceed just like the example used earlier in Figures 8-22 and 8-23.

Third-order equations that can not be handled by the foregoing cancellation


method must be worked out using other means. These other methods can be
extremely involved and time-consuming. We discuss one approach here that will
require a computer or hand calculator that has the programming capability of
finding at least one real root of a cubic equation. We first describe this method,
and then at the end of this discussion we go through a step-by-step set of instruc¬
tions to follow. Some of the concepts presented here are presented without proof
because of their complexity.
The root locus behavior for third-order characteristic equations is somewhat
similar to that of second-order equations. When Kx attains values that cause
imaginary roots to be generated within the characteristic equation, the root locus
leaves the horizontal <x or real axis and begins traveling either up or down through
the second and third quadrants (assuming, of course, that the system is not abso¬
lutely unstable). With the second-order equations, we found that the root locus
left the cr-axis at a 90° or -90° angle, depending on whether the locus was traveling
up or down from the axis. With third-order equations, however, this angle of root
locus departure cannot be depended on to always be at a right angle with the real
axis. Fortunately, we can calculate this angle very easily.

angle of root __± 180°_


locus departure (number of poles) - (number of zeros)

Equation (8-40) works for second-order as well as third-order equations. Looking


at Eq. (8-37), we see there are two poles (5 = 0 and 5 = 5) and no zeros.
According to Eq. (8-40), the angle of root locus departure = ±1807(2 - 0) =
±90°. This is what we found when we constructed the root locus in Figure 8-22.
With the second-order equation, we found that the root locus departed from
the real axis exactly mid-way between the poles of the open-loop equation. This
is no longer true for the third-order equation. Fortunately again, we can calculate
approximately where the root locus leaves the real axis using Eq. (8-41):
248 Chapter 8 Transient Analysis

asymptote
location _ sum of pole values — sum of zero values „ .
point, or total number of poles - total number of zeros
centroid

Notice that we said we can calculate approximately. Equation (8-41) actually


locates an asymptote along which the root locus travels and gradually approaches
and eventually reaches at infinity. At the departure point on the real axis, the root
locus will be at its greatest distance from this line. Let’s look at an example.
Again, we can use Eq. (8-37) to check the preceding equation. The departure
point (i.e., the mid-way point seen between the two poles in Figure 8-22 for Eq.
8-37) would be calculated as:

(0 + 5) - (0)
2-0
This point was in fact what we found out to be true when we plotted the root locus
for Eq. (8-37).
Let’s now try our hand at calculating the actual departure point for a system
and see how it compares with the results of Eq. (8-41). This is not a practical
exercise and is done here only to show the relative comparison between an actual
result and the results obtained from a much easier method that we can depend on
for being fairly accurate.

EXAMPLE 8-6 -
Plot the root locus near the real axis for the following expression and show how its
departure point compares with the asymptote or centroid location calculated with
Eq. (8-41):

Kx( 1 + s)
KG = (8-42)
5(1 + 105)(1 + 0.015)

Solution:
First, we notice that the two poles are well outside the times two amount of the zero,
-1. So we have no choice but to plot the roots for various values of Kx to see how
the root locus behaves near the real axis. To do this, we rearrange the terms in Eq.
(8-42) so that, when multiplied out, the characteristic equation in the closed-loop
expression will have a 1 for the cubic term instead of a fractional value. This also
makes the zero and pole values easier to see.

(0.1X100) Kx{\ + s)
KG =
(0.1X100) 5(1 + 105)(1 + 0.015)

L this fraction is comprised of the


numbers needed to multiply the
parenthetical terms by in order
to convert them to the form seen
in Eq. (8-43).
Section 8-6 The Root Locus Method 249

\OKx(s + 1)
(8-43)
s(s + 0.1)(j + 100)

Since the Kx term represents a general coefficient, we can combine the 10 with the Kx
in the numerator of Eq. (8-43) to get:

Kx(s + 1)
KG = (8-44)
5(5 + 0.1)Cs + 100)

Now, we proceed with determining the characteristic equation just as we did for the
second-order root locus example earlier:

KG Kx(s + 1)
1 + KG ~ s(s + 0. l)(s + 100) + Kx(s + 1)
=_Kx(s + 1)_
s(s2 + 100.15 + 10) + sKx + Kx
=_Kx(s + 1)_
53 + 100. \s2 + 105 + sKx + Kx
=_KAs + 1)_
(8-45)
53 + 100. l52 + s(Kx + 10) + Kx

Looking at the characteristic equation in Eq. (8-45), we realize the work


involved having to derive the roots for a cubic equation for every value we pick
for Kx. However, by using a hand programmable calculator, or by using a PC, the
chore becomes rather simple. We can use the simple BASIC programs for an
IBM PC in Appendix C. Armed with this program or a similar program, let’s
construct another table (Table 8-4) to keep track of our selected Kx values and the
resultant roots. The technique we use is this: We start off with a very high value
for Kx, say 10,000, and then select a low value, 1,000. If our choice results in real
roots only, we know our pick was too low. The root locus is still on the real axis.
On the other hand, if our selection results in generating imaginary roots, our Kx
value is on the root locus part that has lifted off the real axis already. We then
lower our next selection of Kx. We work our selections back and forth until we
“trap” the locus just as it is lifting off the real axis. We want to find the root value
just at that point. The listings of our Kx values are listed in just the order used in
having the roots determined so you can see how the final result was finally “cap¬
tured.”
As the table shows, the root locus leaves the real axis at a value of approxi¬
mately -49.35 (splitting the difference between -49.16 obtained with Kx = 2546,
and -49.53, the real portion of the imaginary root obtained with Kx = 2547).
However, Eq. (8-41) says that the departure point is:
0.1 + 100-1
= 49.55
3 - 1
where 0.1 + 100 in the numerator = sum of pole values. Figure 8-24 shows the
partial plot of the root locus near the real axis along with the placement of the
250 Chapter 8 Transient Analysis

TABLE 8-4 DETERMINING THE ROOT LOCUS NEAR


THE REAL AXIS FOR s3 + 100.1s2 + s(Kx + 10) + Kx

Kx S\ Sl ^3

10,000 -1.009081 -49.55 + 786.34 -49.55 - 786.34


1,000 -1.111094 -10.13 -88.86
5,000 -1.01852 -49.54 + 749.55 -49.54 - 749.55
3,000 -1.031711 -49.53 + 721.31 -49.53 - 721.31
2,000 -1.049281 -26.14 -72.91
2,500 -1.038587 -42.74 -56.33
2,575 -1.037371 -49.53 + 75.37 -49.53 - 75.37
2,550 -1.037768 -49.53 + 71.97 -49.53 - 71.97
2,545 -1.037849 -48.46 -50.60
2,547 -1.037816 -49.53 + 7.93 -49.53 - 7.93
2,546 -1.037832 -49.16 -49.90

asymptote line. The dimensions have been exaggerated to show the difference
between the locus and the asymptote more clearly. As you can see, our use of Eq.
(8-41) to approximate the point of departure is fairly accurate. Besides, since we
will be dealing with points whose complex coordinates will lift them a consider¬
able distance away from the real axis in order to maintain a stable damping ratio,
this error will be reduced even more. As an example, assume that the equation we
have been working with had a Kx of 5,000. The complex roots associated with this
value are s = -49.54 ± j49.54. The real portion of this root (-49.54) closely
approximates the location of the asymptote. Figure 8-24 shows that with this
value, the damping ratio, z, would have a value of 0.707.
Approximating the locus departure point, often called the breakaway point,
using Eq. (8-41) is valid only if the asymptote is vertical to the real axis. That is, if
using Eq. (8-40) to calculate the angle of root locus departure results in a ±90°
value, then we can use our approximation. Any other angle value will demand
some other approach, which we do not cover in this discussion. A typical transfer
function that will not generate a vertical asymptote is of the form:

^l + ST ])(1 + STi)

More advanced textbooks must be consulted for handling this type of case.1
There are other root loci portions associated with a cubic equation transfer
function. There are loci that exist between the various pole and zero locations on
the real axis; this is where you will find the real root, S\, in Table 8-4. However,
we are only interested in the vertical branch root locus, since that is the only
portion of the entire root locus we use for determining system stability.

1 Richard C. Dorf, Modern Control Systems, 3rd ed. (Reading, Mass.: Addison-Wesley Pub¬
lishing Co., 1980), pp. 170-84.
Section 8-6 The Root Locus Method 251

• Kx = 5000 49.54 -

• Kx = 2575

Figure 8-24 The root locus for the characteristic equation s3 + 100. Is2 + s(Kx +
10) + Kx near the real axis.

Let’s list the various steps we took in finding the root locus of a third order
equation:
/

1. Identify the open-loop transfer function.


2. Convert to a closed-loop transfer function using KG/( 1 + KG).
3. Identify the characteristic equation (the third-order equation in the denomi¬
nator).
252 Chapter 8 Transient Analysis

4. Convert the parenthetical terms in the closed-loop expression from the (1 +


5t) form to the (5 + 1/r) form. Remember to multiply the Kx constant in the
numerator by the appropriate conversion factors needed to perform the
parenthetical conversions. (Refer to Eq. (8-43) to review this step if neces¬
sary.)
5. Calculate the roots of the characteristic equation using whatever means you
have available for determining third-order roots. (Refer to Appendix C.)
6. Use Eq. (8-41) to determine the asymptote location on the real axis ( - <j

axis). The location point should be quite close to the real-number compo¬
nent of the complex roots obtained in step 5. Equation (8-41) is used for a
check of the root results obtained in step 5.
7. We can now calculate the following information:

a. Damped frequency: = the imaginary-number component,


or j-value, of the root found in the step 5.

b. Natural frequency: (on = + x] (8-46)

where xr = the real-number component value taken from the complex


root

c. Damping ratio: z = cos(arctan (o„/xr) (8-47)

8-7 OTHER CONTROLLER TYPES

In reality, there are three kinds of automatic control systems. These are:

1. proportional control,
2. integral control, and
3. derivative control (sometimes called differential control).

We discuss these in the next three sections.

8-7.1 The Proportional Controller

Throughout this text, we assumed that all the control systems we discussed were
of the proportional type. That is, the system’s input signal to the system’s overall
transfer function is proportional to the generated error signal. If we represented
the error signal by e and the input signal to the transfer function by E, the relation¬
ship between the two could be expressed as:

K = - (8-48)
£

where K is a proportionality constant. Up to this point, we have been assuming a


Section 8-7 Other Controller Types 253

K value of 1. In other words, E and e were the same. It’s a matter of viewpoint,
however. You could consider the system’s amplifier gain as K since we are as¬
suming the amplifier to be a proportional device. If we separated the amplifiers
from the main transfer function and then let the amplifiers become our propor¬
tional controller as depicted in Figure 8-25, we then have a picture of a propor-

r 1
! E Overall
-J Controller 1- T ransfer -o\/ Out
1 i Function
L J
Figure 8-25 The “controller” in a
feedback system.

tional controller. But what if we wanted a much quicker speed of response and a
decrease in our positional error at all frequencies in response to the generated
error signal? With a proportional control system, our first response would be to
increase the system’s gain. But increasing the system’s gain would buy us a more
jittery system regardless of operating frequency. We then would resort to a modi¬
fication to the controller box shown in Figure 8-25 to give us the desired response.

8-7.2 The Integral Controller

The purpose of the integral controller is to increase the gain of our system at very
low frequencies. We do this by introducing an op-amp integrator circuit where
our controller box is located in Figure 8-25. The circuit is seen in Figure 8-26. By

Rf Cf

Figure 8-26 An integral control circuit.

juggling the values of resistances, capacitance, and amplifier gain in our circuit,
we can actually obtain a combination proportional-integral controller whose gain
curve would resemble something like that in Figure 8-27. The circuit would give
us very high gain at extremely low frequencies but diminish at the middle frequen¬
cies. This is desirable because the very high gain at the very start of energizing,
say, a servomotor, gives you a very fast response time. But as the transient
Chapter 8 Transient Analysis

\n (Rad/Sec)

Figure 8-27 Typical gain curve for an integral control system.

response sets in, you want low gain to be present in order to decrease over-shoot
as steady state conditions settle in.

8-7.3 The Derivative Controller

Ideally, once steady state conditions have been reached, we would like to have
the system’s gain increase once again to be prepared for the next sudden input
change. This is what the derivative control does. Installing the circuit shown in
Figure 8-28 in our controller box of Figure 8-25, we will have the response just

C, Rf

Figure 8-28 A derivative control


circuit.

described. Figure 8-29 demonstrates this particular system’s gain response


action. Again, like the integral control system, the derivative circuit in Figure
8-28 is actually a combination proportional-derivative control system. It’s possi¬
ble to combine the two systems in order to obtain the increase-decrease-increase
gain characteristics to give an almost ideal system response. The circuits pre¬
sented here are just a few of the many circuits used for refining the desired system
response.
Summary 255

w (Rad/Sec)

Figure 8-29 Typical gain curve for a derivative control system.

SUMMARY

The hand-plotting of the transient response curve for a closed-loop transfer func¬
tion can be somewhat an involved process. However, good representations of a
control system’s behavior can be learned this way. There are programs, though,
for the personal computer that can handle this chore nicely. One such program is
found in Appendix B.
The transient response curve in an automatic control system is comprised of
two subcomponent curves, one of which is a sine expression while the other is a
decaying exponential function (hopefully it is decaying and not growing!). Being
able to recognize these two curves allows one to determine the control system’s
vital response characteristics to a step input signal. Actually, there is a third
component. This is a constant that represents the magnitude of the step input
signal. The transient waveform “wraps” itself around this step which forms the
horizontal axis of its plot.
Some of these vital statistics can be calculated from the Laplace transform of
a closed-loop function “shocked” with a unit step input. The transform allows us
to find the natural resonant frequency, the damped frequency, and the damping
ratio of a system. Once these values are found, we can calculate values such as
settling times, response times, over-shoot, etc. However, this technique is good
only for second-order equations. Third-order equations require an entirely differ¬
ent approach, the root locus method.
Euler’s equation helps us to make the leap from the trigonometric time-
related plane of the transient waveform to the so-called 5-plane. It is the 5-plane
that leads us into the discussion of the root locus concept.
When speaking about poles and zeros, reference is made to the open-loop
equation. The term roots is used in conjunction with the closed-loop equation as
256 Chapter 8 Transient Analysis

is the term, characteristic equation. The poles and zeros of the open-loop expres¬
sion determine the shape and direction of travel of the root locus plotted on an 5-
plane coordinate grid as the Kx value is changed in the closed-loop expression.
First-order characteristic equations are the simplest of all equations to deal
with in automatic control systems. However, they are rarely encountered and are
used only as a stepping stone to the more complex equations.
Third-order equations may have their stability characteristics determined
using the Routh-Hurwitz method. The root locus method may then be applied to
find o)n, ajd, and z. Once these quantities are found, all the other transient quanti¬
ties may be calculated. The processes used here on the third-order functions may
also be applied to the second-order functions if desired.

EXERCISES

8-1. Determine the poles and zeros in the following expressions and determine
each equations’ stability criterion based on the determined poles.

25(1 + 0.2s) 65.6(1 + 35)


a. d.
5(1 + 0.3)(1 - 35) 5(1 + 0.55)(1 + 105)

4.87(1 + 6s) 62.4(1 + 0,25s)


b. e.
(1 + 35)0 - 52) (1 + 0.995)0 + 25s)

1044 1,000(1 - 0.25)


c. f.
(1 + 0.l5)(l + 0.5s) 5(1 - 0.55)0 + 6s)

8-2. Plot the transient curve of the function:

5(1+0.155)

8-3. For the expression, KG — 100/5(1 + 0.255), plot its transient response
curve.
8-4. Reduce the Kx value in the expression given in Exercise 8-3 to the value, 5.
Again, plot the transient response curve. What changed in this new curve?
Did the transient curve become more stable or less stable?
8-5. In the revised equation of Exercise 8-4, calculate the following:

a. the natural resonant frequency,


b. cland
c. the maximum over-shoot.
Summary 257

8-6. Plot the root locus for the expression:

Kx
KG
(1 + 8j)(1 + Is)

Determine a>n, cod, and z all for a Kx value of 31.


8-7. Find the ±5% settling time, the number of oscillations to within ±5% set¬
tling, and (om for the expression:

100
KG =
^(l + O.I65)

8-8. Find the damping factor for the expression:

250(1 + 0.35)
KG
(1 + 10j)(1 + 0.275)(1 + 0.95)

8-9. Plot the root locus for the expression:

Kx( 1 + 115)
KG =
5(1 + 85)(1 + 0.025)

Compare your results to the results obtained using Eq. 8-41 for finding the
asymptote location point.
8-10. Using the Routh-Hurwitz Stability Test, find the stability characteristics for
the following expressions:

30
a> O.lj-2 + 5 + 30
h 100
0.252 + 35 - 100
c. Characteristic equation: 53 + 52 + 35 + 12
d. Characteristic equation: 53 — 252 + 45 -8
e. Characteristic equation: 53 + 1452 + 65 + 22
f. Characteristic equation: 53 + 25 - 7

REFERENCES

DeRoy, Benjamin E., Automatic Control Theory, New York, N.Y.: John Wiley
& Sons, 1966.
Sante, Daniel P., Automatic Control System Technology, Englewood Cliffs,
N.J.: Prentice-Flail, Inc., 1980.
9
Designing A Control
System

9-1 INTRODUCTION

The purpose of this chapter is to demonstrate how to apply all the facts that have
been presented so far in designing and analyzing automatic control systems. In
many cases, there is more than one method that can be used in this designing and
analyzing process. Where possible, we discuss the other methods that can also be
used. Sometimes it’s a good idea to have more than one method available so that
one can be used to check the results obtained with the other. And in the designing
process, quite often the first design chosen turns out to be the most expensive
design; therefore, alternative approaches must be used. So we discuss design
examples for various automatic control systems where we can apply our newly
acquired knowledge of system designing. We also want to be able to analyze
these systems and make forecasts of their behavior resulting from making adjust¬
ments to them. The design approaches presented in these examples have been
simplified to keep their developments shortened. However, the overall ap¬
proaches to the solutions are realistic, and it’s these approaches that should be
studied.

9-2 EXAMPLE 1: A ROBOTIC CART STEERING SYSTEM

The first automatic control system that we design is a control system for the
automatic steering of a robotic cart. Here are the details:

258
Section 9-2 Example 1: A Robotic Cart Steering System 259

1. The cart is a self-propelled, all-electric, battery-operated cart (27 VDC, or


two 13.5 VDC batteries in series) that is used for distributing and delivering
interplant mail, maintenance items, and specialized inventory parts for a
large manufacturing plant. You might think of this system as a robotic gofer.
2. Our cart will be designed to follow an invisible infrared reflective paint track
that will be painted on the plant floor connecting the various pick-up points
the cart will pass along its way.
3. A set of photocells will detect the reflection of an infrared light source
mounted underneath the cart, thereby being able to detect the cart’s relative
position to this paint line.
4. The cart’s steering must be designed to allow the cart to stay on its track
despite side forces due to receiving accidental bumps or running over minor
obstacles in its path. The desired damping factor for the cart’s steering must
be in the range of 0.5 to 0.6. This is a good compromise between a not-too-
sluggish system and one that is not too jittery.

Now that we have laid down some basic ground rules for our design project,
let’s see how we tackle it. A good place to start is with laying out a block diagram
of the system. We can fill in the detailed numerical data later. Figure 9-1(a)
shows our initial concept of how we think the steering mechanism will function in
a control system. Figure 9-l(a) is the schematic, while Figure 9-l(b) is the block
diagram. Once we have the design outlined the way we want it, then we can refine

(a) Schematic

Input

(b) Block Diagram

Figure 9-1 The initial design for the robotic gofer cart.
260 Chapter 9 Designing a Control System

our drawings into detailed mechanical and electronic schematic layouts. Here is
how we hope the system will work:
In our sketch we show the photocells that we plan to stride the painted strip
on the floor. We will mount the cells on the cart’s bottom along with the IR
(infrared) light source. The IR source will provide the light energy that will reflect
off the paint into either or both photo-cells. We decided to use photoresistive
cells because they are inexpensive and have good IR response. The cell outputs
will then go to a differential op amp whose output will be either positive or
negative depending on which cell has the stronger response output. The cells’
output of course is determined by how close it is to the painted strip. The error
signal is amplified by a preamplifier and power amplifier. The output of the power
amplifier will then drive a servomotor whose output is attached to a gear box. The
gear box drives a pulley system which, in turn, rotates the steering axle toward the
right or left.
Now let’s see how the system will work: If the cart is moving down the
center of the track, photocells 1 and 2 will be an equal distance from the reflective
strip. Consequently, the signals coming from both cells will be equal. The output
of the op amp, on the other hand, will be zero since its output is the difference
between the two inputs. Since the output is zero (i.e., the error signal is zero), the
servomotor will remain stationary and the cart will continue moving straight
ahead. So far, we see no problems here. Now, let’s see what happens when the
cart is forced to move left due to an outside force such as an accidental bump
received by a passing work person. This means that photocell 2 will now receive a
stronger reflective signal than photocell 1. An error signal is now produced by the
op amp because of the upsetting of the exact signal balance that existed up to this
point between the two photocell outputs. The two amplifiers will now amplify this
new error signal causing the servomotor to rotate, which in turn will cause the
wheels to pivot right. When this happens, photocell 2 increases its output signal
as the reflective strip comes closer, thereby reducing the error signal output from
the op amp. The servomotor then slows down or even reverses if the error signal
goes beyond zero, changing its polarity. If this happens, the steering swings back
toward the left and the process starts all over again.
Obviously, we have to be careful here. We don’t want a system that over¬
corrects or hunts. In other words, we want a system that hunts just enough to get
us back on track with the minimum of oscillations in our steering. Selecting a
damping factor of 0.5 or 0.6 should, theoretically, do this for us. And the very first
consideration we must be aware of is that we want to make sure we have negative
feedback and not positive feedback. Positive feedback would cause our steering
to correct in the wrong direction. A simple oversight such as inadvertently re¬
versing the wiring to our servomotor would do just that. Or, the same effect
would be created if the wiring to the op amp’s signal input would be accidentally
reversed. So we want to check to make sure our prototype design is wired prop¬
erly before operating.
Section 9-2 Example 1: A Robotic Cart Steering System 261

9-2.1 The System’s Dead Band

Another problem that we haven’t discussed up to this point is the problem of a


system’s dead band. Dead band may be thought of as system “stickage.” In our
steering design, the servomotor, along with the rest of the steering control system,
will exhibit a certain amount of dead band because of the low input command
voltages being supplied to it. Even though servomotors are designed to operate at
relatively high input voltages such as ±26 VDC or 28 VAC, most of their operat¬
ing lives are spent responding to input voltages that are just a fraction of their
rated operating voltage. This is because the error signals sent to the servo are
usually quite small. An automatic control system spends a good portion of its
time generating these small touch-up corrective signals to keep the overall system
on proper course. As a result, the servomotor and the components downstream
from the servo’s output must be able to respond to these signals. Obviously, there
is a threshold limit to any servo’s ability to respond to weak signals. This is where
dead band comes into play. Dead band may be defined as a range of input signals
that is furnished to a device or system before that system finally responds to this
input (see Figure 9-2). In Figure 9-2, rotation of this particular servomotor begins
only after receiving a voltage of at least ±1.5 VDC. Dead band is usually ex¬
pressed in angular displacement of the control generating the control voltage to

6 (Rotation)

Figure 9-2 A servomotor’s dead band.


262 Chapter 9 Designing a Control System

the system. In the case of the servo control characteristics depicted in Figure 9-2,
if a rotation of ±5.7° of a control device is necessary to produce the ±1.5 V for
servo rotation, then the dead band is specified as ±5.7°. Dead band is created by
system friction, inertia, magnetic coupling, gear train load, and backlash on the
servo or any combination of these quantities. Any good servomotor should be
able to respond to a voltage as low as about 5% of its rated operating voltage.
Dead band may be calculated using the following equation:
2V
dead band (°) = (9-1)

where Vms = motor starting voltage (volts)


Ks = control device sensitivity (volts/deg)
K — remaining system gain or sensitivity

Notice that only the system’s gain or sensitivity constants between and including
the control device (that is, the pot, synchro, or other device that, when rotated or
moved, creates the activation voltage for the servo) and the servomotor are used
in Eq. (9-1). The servo’s Kx is not included in the calculation.
Dead band may also be expressed as a linear displacement rather than an
angular displacement of a control input. This is explained later as we become
more familiar with our steering system.
One of the easiest methods in making an otherwise unstable system stable is
to reduce the system’s gain. Unfortunately, the price one pays for this decep¬
tively simple solution is an increase in that system’s dead band. You can verify
this by inspecting Eq. (9-1). The dead band varies in inverse proportion to the
system gain. As a consequence, if we decide to lower the system’s gain, we want
to be sure to check out the dead band first to make sure its increase can be
tolerated.

9-2.2 Selecting the Components and Generating the Open-loop


Transfer Function

The heart of our cart’s steering system will undoubtedly be the servomotor. We
want to be sure that it not only has the proper time constants and velocity con¬
stant, but that it also has enough torque to do the steering for us. It’s unlikely that
we will be able to find a servo with a built-in gear train that can generate the
needed torque. We may have to purchase an additional gear box to operate off of
the servo’s already geared-down output in order to create the torque that we
need. We’ll have to keep this in mind when we size the servomotor for the
steering requirements.
Determining the steering force for our cart is relatively easy. We need to be
able to measure the amount of force required to rotate the front axle wheels. This
can be measured by using a calibrated fish scale attached to one end of the steering
cable that slips over the pulley attached to the gear box. The amount of pulling on
Section 9-2 Example 1: A Robotic Cart Steering System 263

the scale that causes the proper steerage will be the force we will use in calculating
the needed servo torque. Let’s assume this has been done and we found that a
force of 3.2 lb. was sufficient to give us the proper steering action. To calculate
the torque, we measure the gear box’s pulley over which the steering cable is
wound, and then multiply its radius times 3.2 lb. Let’s assume a radius of
3 inches; the torque will be 3.2 x 3 or 9.6 lb.-inches. Since this torque amount is
not continuous but only intermittent, we should keep this in mind when we size
our servo. This will help keep the cart’s cost down by allowing us to use a smaller
servo.
At this point, we can start looking for DC servomotors with 9.6 lb.-in.
capabilities, and if we’re successful, we can eliminate the gear box. If we can’t
find a servo with that kind of torque output, then we will have to shop around for a
gear box to augment the gear train on the servo that we do decide eventually to
use.
Along with searching for the proper torque output, we want to make sure
that our torque is outputted at an acceptable speed. Obviously, we can generate
as much torque as we want with the smallest servomotor available by simply using
the highest gear ratio available for a gear train. But the output shaft’s rotation
would be barely detectable. We must decide on an acceptable speed for our 9.6
lb.-in. torque. If we moved the steering cable through a 6-inch travel in one
second, that would give us an acceptable turning velocity. This means that our
drive pulley (the 3-inch radius pulley) would have to rotate with an angular veloc¬
ity of 6 in./sec -r- 3 in. = 2 rad/sec. Since one radian = 57.3°, then the number of
revolutions per second would be

2 rad/sec x 57.3°/360° = 0.318 rev/sec, or 19.1 rpm

We should mention one additional item here concerning the physical size of
servomotors: The body sizes or diameters of the motor housings are cataloged in
whole-number sizes. The values of these numbers are the housings’ diameters in
tenths of inches. For example, an 08-sized servomotor is one having a motor
frame diameter of approximately 0.8 inch. A size 18 housing is one that is approx¬
imately 1.8 inches in diameter.
A likely candidate for our servomotor is described in the specification sheet
in Figure 9.3. Here we see a DC servo that has a peak torque capacity of 163 oz.-
in. or 10.2 lb.-in.1 At first glance, these specifications seem to meet our require¬
ments nicely. Since this torque has been specified as peak, we can assume that
this is the servo’s stall torque (i.e., its torque output at 0 rpm), since the stall
torque is a servo’s maximum developed torque. However, notice that the rpm
appears to be quite high, 573 rpm. Since there are no torque-speed curves accom¬
panying these specs, let’s draw a simple one for our own use from the information
that’s given. Refer to Figure 9-4. We patterned the shape of our curve after the

1 This catalog data is now obsolete, according to the Singer Co., and is presented here only as
an example.
CONTINUOUS
ROTATION
DC TORQUERS

GENERAL CHARACTERISTICS
CM3 6631 003, 004, 005, & 006

Peak Torque (oz*in) 163


Power @ Peak Torque @ 25°C (watts) 77
Continuous Torque @ 25°C (oz*in) 100
Power @ Continuous Torque @ 25°C (watts) 29.5
No Load Speed (r/min) 573
Friction Torque (oz*in) 2.5
Ripple Torque (avg. to peak) (%) 5
Maximum Winding Temperature (°C) 165
Moment of Inertia (oz»in*s2) 42 x 103
Damping Factor (oz*in/rad/s) 2.5
Electrical Time Constant (ms) 1.6
Mechanical Time Constant (ms) 17
Motor Constant @ 25°C (oz*in/Vwatts) 18.6
Theoretical Acceleration (rad/s2) 3.7 x 103
Weight (oz) 21

CM3 6631 003, 004, 005, and 006


COLOR CODED
LEAD 12" MIN.
^ LENGTH 309 MAX

DIRECTION OF ROTATION
CW WITH +ON RED
-ON BK

COLOR CODED
LEAD 12" MIN
LENGTH
SECTION A-A

SPECIFIC WINDING CONSTANT CHARACTERISTICS


CM3 6631
PART NUMBER
003 004 005 006
Sensitivity (oz*in/amp) 59.5 26.2 119 238
Resistance @ 25°C (ohms) 10.2 1.98 40.8 163
Voltage @ Peak Torque @ 25°C (volts) 28 12.4 56 112
Current @ Peak Torque (amp) 2.74 6.23 1.37 0.685
Back EMF (mV/rad/s) 420 185 840 1680
Inductance (mH) 16.3 3.2 65 261

Figure 9-3 Catalog data sheet. (Courtesy of The Singer Company, Kearfott Guidance and
Navigation Division, Little Falls, N.J.)

264
Section 9-2 Example 1: A Robotic Cart Steering System 265

rpm

Figure 9-4 An estimated torque versus


Torque Lb.-in.)
rpm curve.

one in Figure 4-30. Unfortunately, we have no data available for the lower control
voltages as we see in Figure 4-30. The problem we have here for the lower control
voltage curves is, we don’t know how to space them from the 28 VDC curve. We
only know that they will be roughly parallel with our known curve. We also know
that for much lower rpm speeds, which is where we will be operating most of the
time, the output torque will be substantially reduced. We have to keep in mind
that whenever corrective signals are sent to the servo, these signals will be quite
low in voltage since they are momentary in nature and won’t have the time to
build up to full operating voltage. This means that we will be operating quite low
on whatever torque versus rpm curve we have to work from. In addition, we will
most likely have to size a gear box for the servo in order to build up the torque
capabilities once again.
Since it wasn’t specified what kind of servo it is we are to use here, we will
select a viscous-damped motor. Therefore, its transfer function is of the form,
0OU[IEin = Kjs{ 1 + ts). (See Table 4-2.) Our next job is to determine a value for
Kv, the servomotor’s velocity constant. According to Eq. (4-20) in Chapter 4,

where Kv! — catalog velocity constant (rad/volt)


OJf = No-load or freewheeling speed at 100% of rated control voltage

(rad/sec)
100% of rated control voltage (volts)

We could guess at the value of coy at 25% of rated control voltage using our rather
266 Chapter 9 Designing a Control System

crude torque versus rpm curves in Figure 9-4. But that is what it would be—a
crude guess. (A call to the manufacturer will usually get the answers you need,
but not always.) Since this 25% data is entirely missing, our data would be consid¬
erably more accurate if we used the 100% rated voltage data instead and then
derated our results using the methods outlined in Section 4-11 of Chapter 4.
Therefore,

f _ (573)(2tt) rad
(28)(60) volt

27t/60 converts the


rpm of 573 to rad/sec
= 2.143 rad/volt

Since this is data for Kv\ (i.e., catalog data), we use Eq. (4-24) to convert Kv\ to
“real” data. That is, Kv = 2 x Ku\, or 2.143 x 2 = 4.286 rad/volt.
According to the catalog data, the mechanical time constant for our servo is
r ! = 0.017 sec. Therefore, according to Eq. (4-23), r = 2 x r!, or r = 2 x 0.017 =
0.034 sec. Notice that an electrical time constant of 0.0016 sec is also given, but
because of its very small value, it can be neglected. The reason for this is because
of the location of the corner frequency for this particular time constant. The
corner frequency would be located at 625 rad/sec on a Bode plot, far to the right
on the plot compared to the ones that we have done so far. Therefore, the 0.0016
sec time constant would really not contribute significantly to the overall gain or
phase curve. It would contribute if this system were a high-gain, extremely fast
response type system, but this is not the case with our steering system. The
transfer function for our servo now becomes:

6out _ 4.286
^VDSM (9-2)
~E~ ~ s( 1 + 0.0345)

Now, let’s look at the other components of our system. The two photocells
will have a transfer function of ^photo/(l + sr). (See item 10 in Table 4-2.) In order
to keep the response time to a minimum for our light detection system, we pick a
phototransistor for our photocell. These are known for their very fast response
times, typically less than a microsecond or two. We want to make sure that our
phototransistor has a good spectral response in the infrared region, so we want to
be certain this is specified clearly in the photocell’s catalog data. Another reason
for choosing the phototransistor is that it gives us the additional advantage of
possessing a fairly high sensitivity of gain factor. For our transfer function, we
assume r to be negligible and will not be concerned with a value for /Cphoto at this
time. We could determine its exact value now, but it will be easier to compensate
for whatever value it turns out being by simply adjusting, instead, the gains of the
various amplifiers within our system. Or, if necessary, if the gain is too high, we
can install a pot. In most cases, this will be true for many transducers. And to a
certain extent, this is also true in the selection of a gear train. These components
Section 9-2 Example 1: A Robotic Cart Steering System 267

are multiplied together, along with the servo’s Kv, to form the product, Kx, in the
numerator of the system’s final transfer function. Of course, if we knew these
constant values at this point, we would use them. In the case of the phototransis¬
tor and our system design, we can go ahead and order the devices knowing that we
can make system adjustments later with whatever units we have.
This brings us to the next stage of developing our system. We are ready now
to determine the Kx value in our system. First, though, let’s look at a block
diagram of our steering system (Figure 9-5) and its transfer function. The feed-

Figure 9-5 Block diagram for robotic gofer cart steering system.

back path is through the painted strip and its relative positive with the photocell
system. This is shown as simply a line connecting the output of the steering
system back to its input. The transfer function for our system is now:

(^photoX^ampl)(^Crnp2)(^Cmp3)(-^gear)(4-286)
(9-3)
s( 1 + 0.034x)

K.
KC — (9-4)
5(1 + 0.0345)

9-2.3 Generating the Closed-loop Data

We leave our transfer function in this form so that we can go now to a root locus
analysis. We use this analysis to determine the value of Kx needed to produce the
desired damping factor.
Figure 9-6 is a plot of our root locus using the process outlined in Section 8-6
of Chapter 8 for second-order equations. We draw our Kx on a vertical line mid¬
way between the two poles of 0 and -29.41 on the real axis. Since we want a
damping factor of 0.5 to 0.6 (let’s split the difference and say 0.55), we can show
angle 6 as arccosine 0.55 = 56.63°. We can now determine the 7-value produced
by Kx, which is the same as This will be 14.71 x tan 56.63 = 22.33 rad/sec.
To find con, we let ajn = V22.332 + 14.712, which is 26.74 rad/sec. To find Kx,
we must first go back to Eq. (8-5) in Chapter 8. Notice that before the calculating
of (bn, we manipulated our closed-loop expression so that it looked like Eq. (8-5).
We did this by performing the following steps:

1. Convert the open-loop expression into a closed-loop expression, writing that


expression in the 5 form rather than the jco form.
268 Chapter 9 Designing a Control System

Figure 9-6 Root locus plot for KG =


Kx/s( 1 + 0.0345).

2. Then we divided both the numerator and denominator by Kx and multiplied


the denominator by a factor/factor to reduce the s2 term to 1. (The
factor/factor equals 1).
3. We then found ajn by taking the square root of the term under the s2 term,
according to Eq. (8-5).

Let’s do the same thing now for our expression, except in our case we know
o)n already. What we have to find is Kx instead. Therefore:

1. Our closed-loop expression becomes:

Kr K,
s( 1 + 0.0345) + Kx 0.03452 + s + Kx

1
0.03452
+ — +
Kr Kx K?
1
0.03452 29.41 5 1 (9-5a)
x + — + -
K, 29.41 K, 1
1
+ + 1
29A\Kr K

<on = V(29.41 )(KX) (9-5b)


Section 9-2 Example 1: A Robotic Cart Steering System 269

Or,
26.74 = V(29.41 )(KX)

Solving for Kx:

26.742 = 29.41 Kx
26.742
Kx =
29.41
24.31

Looking at Eq. (9-5b), we see that the value, 29.41, under the radical sign, is the
reciprocal of the time constant, r, which is 0.034 sec. We can now make the
general statement that:

(9-6)

or,

Kx = w2t (9-7)
Now that we have found a value for Kx, we have to determine our KgtM. But
first, let’s check our damping factor in Eq. (9-5) to make sure nothing has
changed. This will also check the validity of our Kx value:
From Eqs. (9-5) and (8-7):

s (0.041) (coj = ?
(24.31) (0.041) (gjJ ZS

This fraction removes the 24.31 allow¬


ing the forcing-in of ajn according to Eq.
(8-5)

Solving now for z, we get


2z = (0.041)(26.74)

or,
z = 0.55

We see now that our Kx value checks out and z has not changed.
Now we are ready to proceed with finding Kgear. Since we estimated the
operating speed of our servo to be approximately 150 rpm, and we determined
earlier that we needed an input rpm to our drive pulley of 19.1, we will need a gear
reduction box having a gear speed ratio of rpmm/rpmout, or 150/19.1, or 7.85.
(Since 1 rpm is equivalent to hr radians per minute, rpm is nothing more than an
equivalent form of angular velocity. Consequently, since we already know that a
gear transmission’s speed ratio, N, is cojcoout from inverting Eq. (4-30), we can
also state that a gear transmission’s speed ratio = rpm jrpmout. Since Kgtar =
270 Chapter 9 Designing a Control System

1 /N, our resultant transfer function will be 0.127 (see Section 4-12 in Chapter 4).
If we can achieve this particular reduction, then we will increase our torque
output by the same factor, that is, 7.85 times. With this information, our Kx can
now be calculated as:

Kx = 24.31 = (4.286)CCphoto)(^ampl)(^amp2)(/Camp3)(0.127)

24.31
then (■^photo)(-^ampl)(^amp2)(^amp3) (0 127)(4 286) 44-.7 (9-8)

What this figure of 44.7 represents is the amount of combined amplifier gain
needed, times Kphot05 to stabilize our steering system.
Theoretically, we wouldn’t need to perform any response curves for our
system, since we can calculate virtually every feature of these curves. However,
there are certain things we wish to avoid that are mathematically difficult to
predict. Figure 9-7 is a good example. Figure 9-7 is the open-loop Bode plot of

w(Rad/Sec)
Figure 9-7 Bode plot for 24.3l/s(l + 0.034s).

our steering control system up to this point. Everything looks normal; the phase
margin is right where we would like it to be. But notice where the zero gain
crossover point occurs—right at the corner frequency or bend in our gain curve.
This is coincidental, and not too desirable. The reason is this: You always have
better system stability operating within the -20 dBA portion of a gain curve rather
than the -40 dBA portion. The phase margin changes more slowly for a change in
Section 9-2 Example 1: A Robotic Cart Steering System 271

operating frequency in the —20 dBA area of the curve. It would be nice if we
could move the zero gain crossover point to the left of its present position without
reducing our phase margin too much. Let’s investigate this possibility. If we
were to lower the gain of our system, our gain curve would lower, causing the gain
crossover point to shift toward the left. According to our Bode plot, we could
lower our gain by 6 or 8 dB and still maintain a desirable phase margin. At the
same time, we have moved our gain crossover point far enough away from the
corner frequency to assure us stability with our steering control system. But what
about our dead band? In the case of our steering control system, the dead band
would be measured at the photocell pick-up. It would become a linear measure¬
ment of how far the reflective track would be from the photocell to finally cause
the servomotor to start. Our dead band equation would be:

, , , , . 2 x motor starting voltage


dead band (in.) = —— .-tt-...
(Aamp) (photocell sensitivity)
units are
volts/in. (9-9)

To actually determine this dead band distance without making measurements


would be difficult. We will have to be satisfied with just making relative compari¬
sons. In other words, if we were to actually lower our gain by 8 dB as was just
suggested, our dead band would increase by 2.51 times as compared to the original
dead band.
Earlier, we stated that whenever the gain was decreased in a system, the
dead band would increase. Rather than have that happen, where we would wind
up with a system having little or no response for critical small error signal correc¬
tions, let’s look into another alternative. Let’s consider what would happen if we
were to add a rate generator to our system. An important function of a rate
generator is to furnish damping to a servomotor by shifting the gain curve toward
the left in the Bode plot without affecting the system’s dead band. If the system
becomes too deadened as a result of the increased viscous damping created by the
rate generator, one could simply increase the system’s gain to compensate for this
and obtain a decrease in the system’s dead band at the same time.
Figure 9-8 shows a page of specifications for various sized tachometers or
rate generators (either name is used extensively in the industry). The problem is
choosing the right one for the job. Since the rate generator’s output is dependent
on the rpm of the servomoter that it is attached to, we will want to choose a model
that generates enough voltage with the available rpm to equal at least one-half the
servomotor’s control voltage. This should give us a fair degree of control over the
servo’s operating speed with the full output of the rate generator being fed back to
the servo. Most likely what we will do, though, is to feed back only a portion of
the rate generator’s output, just enough to give us the desired damping. This can
be done through a potentiometer. The tachometer with the 20.8 ±3% volts/1,000
r/min. output looks like a good choice to begin with. It’s transfer function, stated
s

Q
H
sc
w
H
w
CCh

u
o
c/}
LU
LU
CO
o

CO
CMO 9608 R9608 CRO 9610 CUO 9611
Part Number TOO TOO TOO CUO 9612 001 CVO 9612 001

o
o
in
024 002 003 002 003

CNJ
CNJ

00
rH
«H

oo
rH
H
H
rH
rH
H
rH
H
Size 12 12 18

272
+1

+1
CNJ

m
»H
vP

sO
‘Output (volts/1000 r/min) 3±5% 2±3% 7±3% 7±1% 19±3% 7±2% 20.8±3% 7±5% 100±5%
SOO SOO SOO SOO SOO S00

rH
Linearity (% to 3600 r/min) 0.05 0.05 0.05 0.2 (0—1000 rpm)

Output Impedance (ohms) 150 150 1500 1700 275 1000 275 275 1000 300 3000

Ripple Voltage

CM
CO
CO
CO
CO
CO
CO
00

CO
1.5 1.5
(above 100 r/min) (%) (min.)

External (Test)


CM
o
o
100k 100k 100k 100k 100k 100k 100k 100k 100k 100k
Load Resistor (ohms)

CM
m

rH
CM
CM
CM
CM
Directional Error (max.) (%) .25 .25 .25 .25
Over —54°C to

O
O

+i
rH
+i
rH

O
O
rH

o
rH
O
rH
+i
rH
+i

+i

+i
.04
+ 100°C (%/°C) (max.)

Over —15°C to +71°C

*
*

O
CO
+1

+i
rH
m
+i
in
(% Variation from

H—
±.15

(.XllAjUSuas
a-injejadiuaj.
25°C value) (max.)

CO
o
o
o
Max. Speed (r/min) 12,000 3,600 12,000 12,000 12,000 12,000 12,000 8,000 12,000 1,200
oeo SO
Friction Torque (in.oz) (max.) 0.20 0.25 0.25 0.25 0.25 0.25 0.25 0.25 0.25
Rotor Moment of

CNJ
CM
r>»
Inertia (gm.cm2) 16.5 16.5 175
OS
Weight (oz) 2.0 2.0 3.0 3.0 2.8 3.0 3.1 4.5 3.0 15.0

rH
o
o

O <=>
+c

cr
o
o>
CD
in
s
CO
O)
CD
CM

Q.
o
_>

O
CO
CD
£
JC
a>
_>

CO
cn c
*k_
co
o* 3
O <D
k_

</f o
"o
O
O
£
c

«2
<D
*o
a>
o
o

>%
CL
cr
ob
00 o
o> "O CN
O) cr

s
O
c O
O CO

s »S

c
O
'5
o
_
>

CO
DO
O'

Q-»—
o x:
-C c
CD CO
•— CO

CD
>>«
^ a3
c—
■ss
l.
o a>
P/co
3 in
o—
® </>
TD

O) CD k-
in —
<D &>
o
Cl> ^

—' CO
Cl in
00 w
C -o
■O 3
cq .E
£^
<D =
IZS O)
00
<D <✓)

> CL
CL
> CL
Z5 2P

o
S
s
00
TT
o
CN

+Z
o
k_

o
o O)
o
O) o
o
k_
c
CO

o a>
in
CO
Q JZ
<D
E
o
>
c

CO
a>
CO
■O

CL

Q-(-
c

o a)
— k_
«5 E
>,"<5
5 CM
a) c
c a>
*- 33
■° w

.£= w
QO
If
CM X
<rO <2

CO LT3

CO
CO
T m/
03 CO

OO 8 03
+1*10
08 o tn <r»
CM
§• OO O
CM
+ I(M ♦I’lO ♦ I'O
o3
• O 5>
<£>
O O IO

CO 1
CO
03 u
li 1
,EWX- jfc. oa
T

Catalog data sheet for DC tachometer. (Courtesy of The Singer Company,


! §§
♦' io
tn
CO X
8 03 <1
5 e'¬
en
OCD via
•*■1 Csi
o" o
0021 + IO oc § I I
sooo- o o ♦■<m
<£> +'ro
a> +,cu
to
0000 + o o
Via 0929
sooo -
0000 + lli-4
X 1
via oooi
5000- J 8
♦ IO
0000 + o
8 to
Via 2901 ♦'CM
soo-
T

Kearfott Guidance and Navigation Division, Little Falls, N.J.)


£00 + 8
via oo2i -
5000-
0000 +
Via 0929-
5000-
0000 +

via ooo i
5000-
0000 +
Via 2901-
500-
£00 +

i*ofs
Figure 9-8

CM

GO
CO
03

273
274 Chapter 9 Designing a Control System

in units of radians and seconds, would be 20.8 volts/( 1,000 x 277")/60 = 0.199 volts/
rad/sec. These are the units needed to make this transfer function compatible
with the others in our system.
Figure 9-9 is a revised block diagram of Figure 9-1 showing our steering
control system. The tacho is shown in the feedback path. We now have to deter-

Input

Figure 9-9 Block diagram of steering system using a tacho.

mine how this is going to alter the transfer function of our system in the way that
we want. The tacho becomes the H in the general closed-loop expression,
KG/(l + KGH). When we first investigated this equation, we agreed to allow H =
1 for all our closed-loop problems. However, in this case, since the tacho is in the
closed-loop circuit, it now becomes equal to H. In other words, Eout/6m = Kvgs =
0.1995 (see reference 15 in Table 4-3 in Summary of Chapter 4). And, the overall
transfer function now becomes:
24.31 24.31
0.03452 + 5 + (0.1995)(24.31) 0.03 452 + 5 + 4.845
24.31 24.31
0.03452 + 5.845 5(0.0345 + 5.84)
Dividing both the numerator and denominator in the preceding expression by
5.84, we get:
4.163
(9-10)
5(0.0065 + 1)

Figure 9-10 shows the Bode plot of Eq.(9-10). Notice how the corner frequency
has been shifted to the right of its original location. Also, the gain crossover point
appears to have been shifted to the left, resulting in a larger phase margin. We can
easily adjust our phase margin now either by adjusting the output pot of our rate
generator or by adjusting the system’s amplifiers. Adjusting the pot will not affect
the system dead band as will adjusting the amplifier gain.
We now have a very reasonable chance of producing a workable robotic cart
steering system. Because of our efforts in calculating system parameters and
plotting the anticipated response curves, we can be reasonably assured of being
Section 9-3 Example 2: Compass-Directed Auto-Pilot Boat Control System 275

w(Rad/Sec)
Figure 9-10 Bode plot for 24.31/5(1 + 0.0345) with shifted gain curve, or 4.163/
5(0.0065 + 1).

able to produce a system with desirable performance characteristics. Realisti¬


cally, there will most likely have to be additional fine-tuning of the various compo¬
nents being used in our steering system once it is built, but at least we now have a
usable system to work with.

9-3 EXAMPLE 2: A COMPASS-DIRECTED AUTO PILOT


CONTROL SYSTEM FOR A BOAT

In this second example, we want to design an automatic rudder control system for
a small boat. This system will enable the ship’s operator to set a particular coarse
at the wheel and then have the rudder automatically report to that position while
being corrected by a compass for any deviations from that position. Figure 9-11 is
a schematic diagram of the proposed system. Again, this will be a DC control
system. It is assumed that the boat has available a semiregulated high-voltage,
high-current DC supply for operating a large DC servomotor.
We begin our design once again with choosing a servomotor. We pick a motor
having a large torque output capability. The one described in Figure 9-12 will do
nicely. We assume that torque measurements have already been determined as to
what is required for rotating the rudder under the highest boat speeds encoun-
276 Chapter 9 Designing a Control System

Command
m

Figure 9-11 Initial design for an autopilot control for a boat.

tered. Additional torque will be available, since we will have to gear down the
high rpm output of this motor. Notice that the motor’s manufacture has already
saved us the trouble of deriving its transfer function. It is, 1.6/[^( 1 + 0.0225)].
However, this is based on catalog data; therefore, the real world transfer function
becomes:

flout = 1.6 X 2 = 3.2


Em 5(1 + 0.0225x 2) 5(1 + 0.0445) K }
We can check the catalog’s Kv\ by noting that it lists the no-load speed as 3,600
rpm, or 3,600 x 277 +- 60, or 377 rad/sec. The listed rated control voltage is 230
volts; therefore, its Kv\ = 377/230 = 1.64 rad/sec/volt. This checks with the
transfer function’s Kv value. This also gives us assurance that the method used in
our first example for assuming a value for the Kv in Section 9-2 was probably
correct.
The real world expression of our transfer function is based on operating at
approximately 25% of the rated control voltage which, in our case, would be 57.5
VDC. Our problem now is to decide what the rpm would most likely be at this
operating voltage, since the catalog data gives no clue as to what we could ex¬
pect. We can only guess that the rpm will vary proportionally to the supplied
voltage (assuming constant supply voltage to the motor’s shunt field winding).
Therefore, at 25% of rated control voltage, the operating rpm will be in the
neighborhood of 25% of 2,700 rpm (we use a rated rpm figure at 0.33 horsepower
(hp) from the catalog’s chart) or 675 rpm. This figure will help us size the proper
gear train needed to drive the boat’s rudder.
Speaking of the boat’s rudder, we now have to determine the amount of
torque required to turn the rudder under worst conditions. Worst conditions
would be experienced while the boat is traveling at maximum speed, say 15 mph,
Section 9-3 Example 2: Compass-Directed Auto-Pilot Boat Control System 211

Characteristics* — FD84-51-1 FD84 Servo Motor

Rated Power Output (hp) 0.6


Duty 1 ntermittent
Armature Rated Voltage (Vdc) 80
Armature Rated Current (amp) 9.4
Armature Resistance (ohms) 1.2
Armature Compensation
Field Resistance (ohms) 1.1
Armature Total Resistance (ohms) 2.3
Armature Total Inductance (mh) 3
Shunt Field Voltage (Vdc) 100
Shunt Field Current (amp) 0.3
Shunt Field Resistance
at 25°C (ohms) 300
No Load Speed (r/min) 7,800
Locked Torque (in’oz) 350
Torque Constant (in*oz/amp) 18
Starting Voltage (Vdc) 4
Armature Back EMF Constant
(v/rad / s) 0.127
Inertia (in'oz2) 15
Theoretical Acceleration
at Stall (rad/s2) 9,000
Friction Coefficient
(in*oz/ rad/s) 0.43
Time Constant (s) 0.09
Frequency Response (Hz) 2 For polarities indicated, rotation shall be CW viewed
from motor commutator end. (CCW viewed from motor
*Data valid when motor operates from Kearfott shaft end.)
FDF1-44-2211-1 generator with If :n 0.15 amp.

0.33 HP

FD87-12-1 Characteristics — FD87-12-1


ARMATURE FIELD
Voltage (Vdc) 230 70
Current (amp)* 13.7 0.47
Resistance (ohms)** 15.4 145
Inductance (mh)* 1.5 —
Torque (ft oz)* 96
No Load Speed (r/min)* 3600
Inertia (in»oz2) 26
Theoretical Acceleration 17,100
at Stall (rad/s2)
This 1/3 horsepower, 3400 r/min, 230 volt
Friction (in»oz/rad/s) 3.1
direct current motor is an open, fan cooled Time Constant (s) 0.022
©m 1 6
unit. Construction features include flange Motor Transfer Function
Vm S(.022S 4- 1)
mounting, ball bearings, die cast aluminum
Weight (lb) 13.5
housing, lubrication per MIL-G-23827, and
1/ A - — LOCKEO amps 0 7 * At Stall
LOCKED TORQUE: 96
number 18 gauge leads. This unit can be op¬ ** Includes armature resistance of approximately 9
ohms and compensating winding of approximately
erated continuously at ambients of —55°C to ±fc J 6 ohms.
+ 55°C wherein the power loss (input/out-
put) does not exceed 330 watts or the total
winding temperature does not exceed 105°C.

■L X 1 X La ONE END ROUND


FEATHER KEY

COMPENSATING

FOR POLARITIES INDICATED


ROTATION SHALL BE CCW
AS VIEWED FROM SHAFT END

Figure 9-12 Data sheet for a DC servomotor. (Courtesy of The Singer Company, Kearfott
Guidance and Navigation Division, Little Falls, N.J.)
278 Chapter 9 Designing a Control System

and turning the rudder suddenly to 90° to the direction of travel. The force created
on the rudder can be approximated by the following equation:

Fdrag = 1.16i’2Asina: (9-12)

where Fdrag = the drag force on rudder (lb.)


v2 = velocity of boat (ft/sec)
A = area of rudder (ft2)
a = angle of rudder to travel line (°)

Assuming a rudder area of 4 square feet, the resultant force would be (1.2)(15 x
5,280/3,600)2 (4) = 2,246 lb., obviously much too great a force to deal with at that
speed. On the other hand, if the boat’s forward speed were, say 5 mph, and the
rudder suddenly turned to 90°, the Fdrag would be only 250 lb. This means that it
will take a 250 lb.-ft. torque to rotate the rudder under this condition. (We as¬
sume that the 250-lb. force acts through the center of the 2 x 2 ft. rudder area,
thereby creating a 250-lb. torque through the rudder’s steering post.) We size our
steering gear transmission for that condition instead. Let’s assume that we can
allow 4 seconds for the rudder to rotate 90°. This means that it will take 4 times 4
seconds, or 16 seconds to rotate 360°. This is a speed of 0.063 rps or 3.75 rpm.
Our gearing ratio will not be 675/3.75 or 180: 1. We now have to determine our
torque constant, KT. This ratio will enable us to determine the output torque of
our servomotor at any operating voltage. It is calculated using the following
method:

stall torque
torque constant, KT = 100% control v^e (9-13)

In our case, the torque constant would be 6 lb.-ft. -s- 230v = 0.026 lb.-ft./volt.
This ratio changes little regardless of the motor’s operating voltage. In our case,
since we estimated our rpm at 57.7 VDC to be 675 rpm, our torque generated at
that speed would be 0.026 lb.-ft. x 57.5 volts = 1.5 lb.-ft. of available torque.
Operating through our 180:1 gear train, the available output torque at the rudder
will now be 180 x 1.5, or 270 lb.-ft. This means that there is enough torque to
turn the rudder a full 90° in four seconds at 5 mph. Our concern now is, will this
slow turning rate create too sluggish of a system for our steering? Also, we will
have to be careful that no attempt is ever made to turn the rudder suddenly to 90°
at boat speeds higher than 5 mph; otherwise the rudder will stall out and not
move. However, there is one consolation in our design: Regardless of how fast
the boat is traveling, the relatively small touch-up course corrections due to wind
and water currents will most likely involve only small rudder movements. This
means that the rudder drag will be kept small, resulting in much faster response
times for our control system. Figure 9-13 shows our control design results so far.
What isn’t shown in the block diagram is the servopot attached to the rudder to
indicate the rotation amount. The pot sends back a voltage proportional to its
Section 9-3 Example 2: Compass-Directed Auto-Pilot Boat Control System 279

Figure 9-13 Autopilot control system after sizing gear train.

rotation and is compared to the difference voltage coming from a voltage compar¬
ator and servopot attached to the boat’s compass.
Our first step in analyzing our initial design results is to plot the root locus
(Figure 9-14). Our system transfer function is:

$out _ ^amp 3.2


(9-14)
Em ~ 180s(l + 0.044s)

where K UWK3-2)
(9-15)
180

Figure 9-14 Root locus for ^amp 3.2/


1805(1 + 0.0445).
280 Chapter 9 Designing a Control System

We identify the open-loop poles and place a locus mid-way through the distance
between them, knowing that whatever Kx value we select, it will lie on this vertical
line. We pick 6 to be the arccosine of 0.6 (assuming we will want a z of 0.6) which
is 53.1°. We can now calculate cod, which is (11.37) (tan 53.1°), or 15.14 rad/sec.
Using Pythagorean’s equation, we find ajn to be 18.93 rad/sec. Using Eq. (9-7), we
can now determine Kx, which we find to be 15.77. And we can now find our
amplifier gain by using Eq. (9-15) and solving for Kx:

(Kamp)(3.2)
15.77
180

(15.77)(180)
Solving for Kamp* ^amp
3.2
= 887

Since rudder response time to a 90° rotation command is a concern, let’s


determine the settling time for the rudder. In Chapter 8, we found the ±5% to
settling time to be 3/zcon. In our case, that figures out to be 0.26 sec. This is small
compared to the four seconds needed for the rudder to arrive at a 90° position. It
appears that this is the only concern that could cause a problem with this particu¬
lar design, aside from the fact that servomotors of the size that we are using in this
design are quite expensive. Now that we have completed an autopilot control
system design for our boat, we study another problem, a design problem for a
remote valve positioner.

9-4 EXAMPLE 3: A REMOTE VALVE POSITIONER

Our third example of an automatic control system uses AC circuitry instead of DC


circuitry just to get you familiar with AC control hardware. This particular pro¬
ject is a remote controlled valve stem positioner. The valve could be part of a
fluid control system in a chemical plant, or perhaps part of a flow control system
for gases or coolants in a nuclear facility. This system is more of an instrument
control system. That is, the hardware is physically small and is capable of low
torque generation only. Supply voltages for this system will be 115 VAC, 60 Hz.
Figure 9-15 shows our design layout. The block diagram for this system is shown
in Figure 9-16. Our set point input to our system is shown as 6in in Figure 9-15.
The control scheme will work like this: The control transmitter’s rotor is the
setpoint input. The amount of rotation, 0in, relative to the angle of the control
transformer’s rotor, determines the phase and magnitude of the AC signal going
into the amplifier. The phase angle is measured relative to the AC supply volt¬
age. The servomotor’s rotor direction and amount of rotation is determined by
the amplifier’s output voltage phase and magnitude. The gear train’s rotational
output is mechanically coupled to the valve’s stem. It is also coupled to the rotor
of the control transformer. If the two rotors are at the same angle of rotation, the
Section 9-4 Example 3: A Remote Valve Positioner 281

Mechanically Coupled
-
115 VAC \

Figure 9-15 Schematic of remote valve positioner.

error signal will be zero. Any difference in these two angles will create an error
signal in the form of an AC voltage, which will cause the servo to turn in the
proper direction to reduce this difference to zero.
Once again, we begin our design project by searching the design catalogs for
the proper servomotor. Because of the relatively light torque requirements re¬
quired to operate the valve, we can probably get by with using about a number 8 or
10 servo. Looking through the specifications, we come across a number 11 servo¬
motor, our only choice according to the data sheet, that operates at 115 VAC 60
Hz. See Figures 9-17(a) and 9-17(b). (Notice that many of the other servos oper¬
ate at 400 Hz according to this sheet. This is because 400 Hz is the line frequency
most commonly encountered in aircraft. This higher frequency is used to reduce
the sizes, and therefore the weights, of power transformers needed for voltage
rectification on board the aircraft.) We can now assemble the transfer constant

Set Point
Input

Figure 9-16 Block diagram of remote valve positioner.


r-s oo rs g*S«—'-ho cn r~-- Lr? oo
S3I83S Z 3SVHd in«3-inSoSoooo 1 : i—i cp a~> cp ■—i cd

(ohms)
»—<«—i i-i ro ^ ro cn co csi csi r-H r^- cm co

*z
NCON«JN°N°rH in in cd cn cd oo
I 3SVHd in<4-ifi^-in3ir)9(D r^rvocnSrHO
1—1 ■—' cn OO 3" CNI t—1
(T)'-' cr)*-* csi csi csj»—i1 " cni cn

w u
o
N co^cn^tD^cooLn S nr CSI CSI «o
i—i (ohms) S3I83S Z 3SVHd csi^csj^rcngcnoocn Pgr^csito^r^r^
r-H CSI CP r-H r—I 'O
c/5 rHCnrH°CJNCgCV,C>g r* H
*X

+
Q O O O IS ,n O oo

2430

2430
122.5
123

123
123
123
oo oo nr nr cd on

94

94
£ I 3SVHd N CD CD [3 CO is
< C
3
o ^co^gaogaooo.—i 2 oo nT nT m
S3I83S Z 3SVHd oo--ioo<f?oo^oonrcD xg csi nr cn oo csj n (33
Cj3r-4CT>JI^«—'CMCSJ ; 7 r-H r—i oo cn r-H CSI Lh
(ohms)

S-H
W 2
*R

O
N o
i—(
00 ^ CD “1 CD ^ ^ g LO CD CD O LH CD O 6
00 oO i—i oo i—i oo sr oo 22 ^3 uo in ^3- g _ rs. oo i/j <u
I 3SVHd csi csi csi 22 52 csi in c/5
p< 00.—IO0.—iCTogj0’^]'—' rH rH rH 01 ^ rH CM O,
<u
O a>
O £ i~
H
Uh CDOGD^.O^OlO'
LO
„ cn cd in i—iais 3
3
03u.
CURRENT

00
O
S3I83S Z 3SVHd coootof^csigcsjo^
.—1 rH H .—1 ^ 1 ,—j nH
cd in oo is cd »—i
^ * rH r 1 i \ | v}' rH rH
03
S (U
U a

H O COOCDOtDroCD rH Q CL)

00 > I 3SVHd CO OO CD OO CD
1—I,—l<—h—1.—1
tO CD
1—1 *n,—1
cn oo o o csj is
m in in o ^ m oo o
o
w o
w oo MOIOVJ 83MOd
co in ro in oo cvj oo in lo
CDNCOr^COrvlDlDCD
nr ^ cd °o cn
E
H n tn ^ n ^ ^ °9
*P0WER

• C/5
U N U 3
< H s ^ a
P4 ('XBlil) (S}}BM) oo oo oo oo ld oo oo

All motors operate continuously at stall.


—:IC—:!C—! CD
W 3SVHd 83d iOdNI 83M0d CO^CDf^COCMCOCSICD cn cn cn cn cro cn cn si *7| 3 OJ
< . -*->
X E
u
N O &o "2
-3 — E
u
o ( UlUi) (LUdj) ooooogogo O CD 2 O
ooooo°og° CD CD S CD
CD
O
CD
CD
CD
CD
(U
f-l GO <3
+-> *-< ■s^
o in LO LO in un CD LO CJ CD csi csic—Lin in in o ^
O <U c
hJ H- G33dS avoi ON CDCOCOOOCDOCDOO CD CD O CO
t—H r-H r—i r-H
CD nr cn
o 73 3 °
c.
< ^ .>
_
DO,
3
h—*
u Q
( U1UJ) (70 Ul)
a) '3 (3 03 *—I
I—I O'tOOOOOOOOOON o o m cd in in o a>
Z CncncnCSlCSICSJcnCSIcsj CD CD OO is ; IS co a
< <
X 3noaoi nvis C/5

a> ^ <u <u


E
E o ■£ <3 u, ■*-’
CJ so 5.1 *->
3 (jQ
oo « ao <=> «*? o tCj <=? cn ^S^oo^aoP zr
(•XBUi) ONIldVIS •.—i ■ co'—1 oof ; >—i ■ °° ,-H - r-H - ,-3 03 3
w *-—» +—» J—<
2Q- 3•—
VOLTAGE

S < £
(volts)

<u
CO CD CD CD CD CD CD CD 1£2 CD CD CD CD CD CD
3 £ Cu
S3I83S Z 3SVHd CMCNjcvjooro^Zico^-cv) "cnsrcncntncn <u o
Q c <*>
3 ' E
£
LO LD LO LO LO LO LO 0 N < 3 3
< CO CD CO CD CD CD Jf3 CD r-H r-H r-H r-H y—K t—H r-H
I 3SVHd CVJ CVJ CNJ CVJ CNJ m CVJ ^ CVJ r—\ r—H »—H »—H r—H f—H r-H a t:
E
n 2 X
03

< ^ S pi s s
CR4 0164 002**
PART NUMBER

U * * M—
CR4 0132 012
CRO 0130 570
CRO 0132 670

CRO 0133 670


CL9 0121 009
CL9 0118 017
CL9 0121 003

I—H *
0£h
H
P121-02

R119-2+
R119-36
P124-06
R118-2

R124-4
R118-1

R124-1

U
w

PJ
ii

2
SIZE

282
bfl E M E CO
O o
cd r"- cd 3 TD
c- CD i--- cd o
i-h CD
O CD o o
O CD
^ o
o o
T-H o
CD CD
X
TO 4= 00 CO
CD

2 <= + r + 1
+ r + i + 1 1
O
CD
E
1 1 1 | I | |
03 CD
>
t-H
03 CD O O CO
CD
o CXI CvJ CvJ CvJ CvJ CD Old T3 ■E?.e bfi
r—l 1—1 r-H t-H T—H T—H CD O 3
CD O 3 M
O CD txo O
>

O cz
CD CO CD CD
CM
CO
CD CD CvJ CN
+ I s
CD TD
CD
M
o o o o o CM t—H O LT> +1 TO ^ C
OO OO OO OO OO T-H OO
<3- T-H r-^ \ \ \ o o CD C Q) ^
CD CD CD CD CD CD LO CD CD
CD LO LO CD O
CD CD INI
_ CO t-H t-H CN CvJ O
GO ° E
+ <d o

o
rH
O
O O
r-H
O
t-H
CO
t-H
f-H m
o o
t—i
CD t-h
CD OO CD OO
CD CD
o
o
CD Mg
cog
CO
CD
o o CvJ
CD O CD o O o oo o o O O CD O O CD CD
o o CD O
J3
(in.;

.373=+=.

. 374=!=.
"O

.363 +
DIM.

+i +i +i + 1 3
+i
CD CvJ
+i
CD CD CD
+!
CD
+ 1 + 1 +i + 1 +i
3
_ o
CD LO LO CO oo CD CO C
rH CO r-H T—H
CD LO t-H r^ r- t—H T-1
r—H \ \ T—H co CN 3
CvJ co co CM O
LO LO IT) UD ■O o
^^
o
Q_ .
CD 4-*
in CO
C\J — •i= CL
CO
3 '—t>0 3 3 3 3 3 3 3 o CL *—
3=
GO
>— u. c
■- T=) 5
M
O

3 -
c
03
-a c:
CD <t
i—
E
to
O

3
O

3
O

3
O

3
O

3
3
O

3
O

3 TO
O

3
3
O

3
3
O

3
CO
o
-g ro JD
E
o
CO ‘ Q_ q_ cd_: Q_
D_ Q_ Q_ Q_ Q_ ol
Q_
a_ Q_ Q_ "o O• 13
<o o
00 c=
O- o
03 CZ
ol
TO CD O k_ * ■"»>
4o—
ID —
4—* - ol 3 CD
LU
o CO CD CD
^1 < 00 o Q LU r-- CO CD o CO o CuO ■o
k_
t—H CD X
GO Ol CN CNI CM CM CN O O 3 CO
o o o E ^ CO “O
oo t—H t—H
E ^ biO c
rr' CD «c CO O o“ LU t-H LU
< Li- CO LU
o
<C^
rH rH t-H rH r-H
CN
t-H
CN
t—H
CO
t—H
LU
t—H
CN
CD
CM
O
CD
O
CD
CD CO bO 4-»
CO

CO
•&
i i
a- =) 00
rH
t-H
OO
T-H
r-H
00
t-H
t-H
00
rH
rH
oo
rH
i-H
CD
O

CD
1
o

CD
i
CN
■^r
CN CN
t—H
r—H t—H ■'Cf- CQ
-5 <° —
00 CD
3

CO
</>
Q

03 O »-
Ol Ol oc. Ol oc O O o CC QZ QZ CL. Q_ Q- Q_ >v
O _ro o'0 §
CD £
a. • >,S £ o *c:
-o
CO 03 CO s:
TD TO
3 o Q.^" si **- —» a

DZ 3 3
txO O «»n i
Ci.

M
£
<*> CO —
b0 -.CN •c 2 5
*3
c to to
3 "2 TO> CO
a) II<D T—*
O biO oz .a +1 J=> cc
"
£

s;
• "»«<

CD
S.2 rH
r—i

co on a
C;
^
.
CN
o k—
CL
o
CO V£j/-
o TD
4— TO
£
<D
% oo O O CD
o CD CD vj- f-H <D 0)
A3N3DIJJ3 rH > 4-* s:
D Sw-
o O O O O C/3
%a0l3V3 O O
CD <D 00 o w 3 4—*
3 TO
M3M0d C33
O c O «
O
3 4-j
«=•£ QZ cd
co E s;
3
CD TO
3 "O
Cs
GO —
O 3 H—* _
o
*3 TO *
4-* 44-»- 3 4-*
s-
of O o
4-»
•j§£
o E
rH
TD -tf 3 CO
3* o
c
C C CD — >
IN — o
t-H
x-P to CO o u
*
CD ID
CD
CD
CL
k_
o CO a
3
o 0/j
00 -Q ^ 4— ID •

§S E ° CD a
CD -f-> bj) CD

CO *=;
CO 3 3 CD
« a
CO O
■n ^

|s. CO C
3
TO r:
§“
> 4-»-
^ O
CD
O 3!
O' ^
sr
CU)g_ | W>° oo
4—*
CD
->—< 3 03 CD
k>
<4j
CD 13 3 CD O O
E ro o
4- TD
O 03 13 13 WD
e
.<5 a> a)
id -c
CD
.E CO
4^
IZ a

*- o lo §
^2 >» CD 4~» 4O
-» —*
CO -Q CO 4-* 4-»
4o- <_) -o
- O) o> 4—
• CO
"O CD ^ 13
4- C </> 3
C JZ
— CO
iT) vT rO oo — o 3 — ^ O —
SiiVM O CL
00 3
co O .E
% O CD oo CO ^ O
GO E ^
O co
3. O -3
ADN3IDI3J3 OO Q- 4—* co <3°

% dOiDVi O o O O O
«3AA0d CD CD <T C\J CNI

283
284 Chapter 9 Designing a Control System

'PARALLEL TUNING
DC
'EFFECTIVE R

CAPACITOR FOR

ROTOR MOMENT OF INERTIA

THEORETICAL ACCELERATION
RESISTANCE

UNITY POWER
(ohms)

FACTOR (Mfd)
(ohms)

TIME CONSTANT (sec.)


(gm cm2)

(rad/sec2)
PHASE 2 SERIES

PHASE 2 SERIES

WEIGHT (oz.)
SIZE PART NUMBER

PHASE 2
PHASE 1

PHASE 1
PHASE 1

SERIES
R118-1 250 250 53 53 2 2 .46 46,000 .0148 1.45
R118-2 190 190 53 53 1.7 1.7 .46 52,000 .0087 1.45
R124-1 250 250 53 53 2 2 .46 46,000 .0148 1.45
R124-4 190 2000 53 584 1.8 .16 .46 44,000 .0116 1.45
10 P124-06
P121-02
CL9 0118 017
250
5280
250
480
5280
479
53
740
53
91
740
91
2
.09
2
.95
.09
1.04
.46
.46
.46
46,000
44,000
46,000
.0148
.024
.0148
1.45
1.7
1.45
CL9 0121 003 4920 559 740 41 .08 .37 .46 43,000 .0244 1.7
CL9 0121 009 248 475 166 37 1.3 .98 .46 43,000 .0240 1.7
R119-2 3800 3800 438 438 .16 .16 1.07 39,450 .0168 4.5
R119-36 3800 372 438 42 .16 1.5 1.07 39,450 .0168 4.5
1 1 CRO 0130 570 4250 476 356 42 .15 1.25 .76 32,600 .032 3
CRO 0132 670 3770 376.2 255 26.5 .175 1.75 .76 69,700 .0113 3
CR4 0132 012 1570 153 105 11.2 .44 4.5 1.2 139,000 .0097 3
CRO 0133 670 3790 372 340 33 .15 1.51 .76 74,000 .0086 3
CR4 0164 002 3745 371 1600 156 .48 4.8 1.07 39,450 .0085 4.5
When phase 2 windings are connected in parallel, the following relationships exist:
Current = 2 x phase 2 series values; *' *2

R, X, Z, Effective R, and DC Resistance = lA phase 2 series values;


Voltage = Vi phase 2 series values;
Parallel Tuning Capacitance for unity power factor = 4 x phase 2 series values. TERMINAL UNITS
*1-18 3, *2 -28 5. 684
LEAD UNITS
*1 - TW 8 WH
CR4 0164 002, CR0 0130 570, *2 - BK 8 GN, RD-BK 8 RO
CRO 0132 670, CR4 0132 012, AND CR0 0133 670 R119
*CR4 0164 002 (which has tapped holes
mounting face) is 1.391 in. (max.) long. j05O4005-
CR4 0132 012 (which has radial leads) is 1.187 093*003
in. (max.) long.

SIZE 10 & 11
<2 0>0<I
OOQ 8§° _, _
000<J

PINION DESIGN DATA ??§


No. of Teeth 13
Diametral Pitch 120 MIN. LENGTH-

Pressure Angle 20°


Standard Pitch
.1083
Diameter (in.)
Tooth Form Full Depth Involute SIZE 11 SERVO MOTOR
CRO 0132 670
AGMA Quality Class Precision 1 04
a.u.
2247 +-0000 1 1 I
Outside Diameter (in.)
i/TORQUE p DWER

Testing Radius (in.) 0546 + *0000


0012
3 rSP EO FA CTQR^

WA TTS
input
Generating Pitch 1
.0538 minimum EF 'ICI ENC Y
Radius (in.) TS
0 UTP UT.
Material AISI 416
Hardness Rockwell C28 to 38
.3 4 5
TORQUE IOZ INI TORQUE IOZ INJ

SIZE 11 SHAFT DIMENSION DATA TYPICAL KEARFOTT DRIVING AMPLIFIERS FOR


PART NUMBER SHAFT TYPE DIM. A (in.) SIZE 10 AND SIZE 11 SERVO MOTORS

R 119-DA Pinion 494 +009 MOTOR PART NO. AMPLIFIER PART NO.
— .006
CL9 0118 017, CL9 0121 009 C70 3148 001
R119-0 D Plain 494 +-009
— .006 C70 3148 OOl
R124-40
R119-0 E Pinion ,408±.010 (transformer required)
CL9 0121 003 P124-06TI C70 3189 001
Rii9-n j Pinion ,384±.009
CRO 0130 570 R119-360, CRO 0132 670 C70 3148 001
CRO 0132 670 Pinion 437±.015 C70 3189 OOl
CRO 0133 670 R119-2D (transformer required)
CR4 0164 002 Pinion ,494±.015 CR4 0164 002 C70 3302 001
CR4 0132 012 Plain .437±.015 CR4 0132 012 C70 3516 001

Literature detailing other driving amplifiers for these units Is


available on request.

Figure 9-17(b) Servomotor data sheet. (Courtesy of The Singer Company,


Kearfott Guidance and Navigation Division, Little Falls, N.J.)
Section 9-4 Example 3: A Remote Valve Positioner 285

from the given data. The time constant, r! = 0.0085 sec. Therefore, r = 2 x
0.0085 sec, or 0.017 sec. The Kv\ value, according to Eq. (4-20), is (3,000)(27r)/
60 -T- 115 = 2.73 rad/sec/volt. Kv = 2 x 2.73 — 5.46 rad/sec/volt. Our transfer
function now becomes:

5.46
KG (9-16)
s{ 1 + 0.0175)

The next piece of hardware that we must select is the control transmitter and
control transformer pair. Since we are using 60 Hz for our supply frequency, we
must also select 60 Hz synchro units. The units that we choose are marked in
Figure 9-18. But notice the sensitivity data given for them. The control transmit¬
ter’s sensitivity is listed as 44 mV/deg, whereas the control transformer’s sensitiv¬
ity is given as 70 mV/deg. Since these units are designed to work with each other,
we should have a mV/deg figure that is given for the pair. Unfortunately, manu¬
facturers rarely do that, so consequently our only choice in this matter is to ignore
the figure given for the control transmitter and use the one given for the control
transformer. We assume that this figure is valid for the two units operating as a
pair. In other words,

Kcx-ct — 70 mV/deg
= 0.070 V/deg
0.0175 rad/deg
= 4 V/rad

Now that we have established transfer functions for our synchro pair and
servomotor, we have to determine the system’s overall transfer function. Since
we don’t know Kamp and l/N, we must go to a root locus plot to determine our Kx
as we have done in the other design examples. Finding that value, we can then
determine the combined value of A^amp and 1 IN. Once that is determined, we can
then make some reasonable guesses as to what our gear train ratio should be.
Going to an 5-plane plot, we lay out our poles as shown in Figure 9-19. We
have a pole occurring at cr = 0 and at cr = -58.82. Our root locus will be located
mid-way between these poles, or at cr = -29.41. Only the upper vertical locus is
drawn. Again, selecting a z of 0.6, angle 6 becomes the arccosine of 0.6, or 53.1°.
Then,

cod = (29.41)(tan 53.1°)


= 39.17 rad/sec

and,
<u„ = Vcorf2 + 29.412
= V39.172 + 29.412
= 48.98 rad/sec
286 Chapter 9 Designing a Control System

SIZE 11-60 AND 400 Hz SYNCHROS

FUNDAMENTAL NULL VOLTAGE (mV) (max.)

MAX. ERROR FROM E.Z. (minutes) (max.)


DC STATOR RESISTANCE ±15% (ohms)
DC ROTOR RESISTANCE ±15% (ohms)

ROTOR MOMENT OF INERTIA (gm cm2)


TOTAL NULL VOLTAGE (mV) (max.)
INPUT POWER (watts) (nom.)

FRICTION @ 25°C (gm cm)


INPUT CURRENT (mA) (max.)

OUTPUT IMPEDANCE (ohms)


INPUT IMPEDANCE (ohms)

SENSITIVITY (mV/degree)
TRANSFORMATION RATIO
OUTPUT VOLTAGE (volts)
INPUT VOLTAGE (volts)

(output open circuit)

(input open circuit)

PHASE SHIFT (°)


PART NUMBER

WEIGHT (oz.)
PRIMARY
TYPE

*
RS911-1 R 26 280 .95 107/81.7° 18.1/79.5° 9 3 11.8 ,454±4% 206 5 26 17 4 2 4

RS911-4 R 26 90 .30 359/81.3° 60/78.5° 28.5 10.5 11.8 .454±4% 206 4.7 26 17 4 2 4
*
RS911-7 R 26 136 .45 236/81° 40/78.5° 19 5.75 11.8 .454±4% 206 4.7 26 17 4 2 4
137 .783±4% 94 * 4 4
RS911-2 R 115 60 .80 2210/82.3° 1130/81.3° 159 90 1570 4 59 2
*
RS911-3 R 115 50 .97 2598/81.8° 60/78.5° 218 10.5 18.2 .158±4% 318 5 42 26 4 2 4
*
CX RS911-5 R 115 30 .44 4670/81.1° 2510/79.3° 320 318 90 .783 ±4% 1570 5 94 59 4 2 4
★ 4
RS911-6 R 115 70 1.03 2060/80.8° 18.1/79.5° 160 3 11.8 .103 ±4% 206 5.6 26 17 4 2
26V 11 CX 4c R 26 130 .37 244/82.3° 41.4/82° — — 11.8 .454±.009 206 4.25 19 12 7 5 2 4.7
11 CX 4e R 115 31 .49 4210/82° 2170/78.2° — — 90 .783±2% 1571 4.5 75 45 7 5 2 4.7
R 195 22+j33 — — 44 * 4 4
R911-03** 6.3 .58 4.8+j5.5 2.5 ,398±4% 31 18 12 2
TX 26V 11 TX 4c R 26 280 1.0 106.1/83.1° 18.5/79.5° — — 11.8 ,454±,009 206 4 — — 7 5 2 4.7
RS941-1 S 11.8 165 .25 74.3/79.7° 86/77.1° 17 10.5 11.8 1.154±4% 206 6 26 17
* 4 2 4
RS941-4 S 11.8 65 .097 195/80.8° 231.2/76.7° 49 21 11.8 1.154±4% 206 7.4 26 17
* 4 2 4
RS941-2 s 90 60 .6 1640/80.7° 1990/77.3° 385 195 90 1.154±4% 1570 4.7 94 59
* 4 2 4
CDX 26V 11 CDX 4c s 11.8 150 .25 76.5/79.5° 88.3/77.3° — — 11.8 1.154±.023 206 6 26 17 7 5 2 4.7
11 CDX 4b s 90 49 .53 1820/82° 2165/78° — — 90 1.154±2% 1571 4 90 60 7 5 2 4.7
R941-03** s 2.4 80 .075 17+j30 37+j35.2 — — 2 .980±4% 35 31 18 12 10 4 2 4
RS901-1 s 11.8 165 .25 74.3/79.7° 418/78.3° 54 10.5 22.5 2.203±4% 393 6 53 34 * 4 2 4
RS901-3 s 11.8 21 .03 577/80.7° 3340/79.2° 385 60 22.5 2.203±4% 393 4.2 40 30
* 4 2 4
RS901-2 s 90 20 .18 5470/80.8° 3340/79.2° 385 555 57.3 .735±4% 1000 4.5 94 59
* 4 2 4
CT 26V 11 CT 4d s 11.8 86 .142 131/79.7° 704/79.8° — — 22.5 2.203 ±.044 393 4.7 18 15 7 5 2 4.7
11 CT 4e s 90 18 .20 5025/80.4° 3370/80.36° — — 57.3 .735±.015 1000 5 60 32 7 5 2 4.7
R901-03** s 2.4 25 .02 49+j90 400+j500 — — 4 1.90±4% 70 26 25 15
* 4 2 4
* Available with 5, 7, or 10 minute accuracies. When ordering, preface basic part number with accuracy required e.g., for a 7 minute RS911-1, order 7RS911-1.
10 minute components require no numerical prefix.
** 60 hertz units. All others have a frequency of 400 Hz•

BuWeps Type Military Part No. Pinion Shaft Data (All Units)1-

; AX 26V 1 lCX4c M20708/8C-001 No. of Teeth 21


llCX4e M20708/2C-001 Diametral Pitch 120
26V llTX4c M20708/6D-001 Pressure Angle (°) 20
t 7C +.000
26V HCDX4c M20708/9C-001 Std. Pitch Dia. (in.) •I/O _ 002

1lCDX4b M20708/81C-001 Max. Root Dia. (in.) .155


1 Q79 + .OOOO
26V llCT4d M20708/7C-001 Outside Dia. (in.) .lo/Z_ 0005
llCT4e M20708/1D-001 Tooth Form Full Depth Involute

+ Shaft length on all units is .555 in. to stop on shaft

TERMINAL
CONFIGURATIONS

RS900
RS9I0
RS930
RS901, RS911, and RS941 units are supplied with leads. Identical units
equipped with terminals are identified by basic part numbers RS900,
Outline dimensions for Bu/Weps units are in accordance with Mll-S-20708. RS910, and RS940 respectively.

Figure 9-18 Synchro data sheet. (Courtesy of The Singer Company, Kearfott
Guidance and Navigation Division, Little Falls, N.J.)
Section 9-4 Example 3: A Remote Valve Positioner 287

Figure 9-19 The root locus for the


valve controller’s transfer function.

We can now calculate Kx using Eq. (9-7):


Kx = on2T
= (48.982)(0.017)
= 40.78

And since: Kx = (KCX-ct)(Kv){K^)(\IN)


40.78 - (4 V/rad)(5.46 rad/V)(^amp)(l/N)
(Kamp)(l/N) = 1.87 (9-17)

Our next step is to select a reasonable gear ratio for our gear train. This
decision is based on how quickly we want our valve to open and close and how
much torque is available. Assuming that very little torque is required for this
particular valve application, let’s assume further that we wish to have the valve’s
stem rotate at a speed of 10 rpm. Assuming an error-correcting rpm for the
servomotor as being 25% of 3,000 rpm, or 750 rpm, we can now calculate our gear
ratio:

Ain 750
Gear ratio, N =
^out 10
75

Then, from Eq. (9-17):

Kamp = (1.87)(75)
= 140.3
288 Chapter 9 Designing a Control System

Often, it is difficult to order the exact gear ratio needed from the manufac¬
turer. You must choose the closest ratio available and then compensate for the
change in the system’s gain by adjusting the amplifier. If compactness is required
in the automatic control system, you can often specify that the gear train be
directly attached to the servomotor’s casing for the purpose of creating one uni¬
fied system.
Since we are dealing with a control knob that is to be rotated to a desired
setpoint (the control knob being attached to the rotor of our CX) for the purpose of
positioning our valve stem, let’s calculate the dead band within the control system
using the recommended gain calculated from Eq. (9-17). We do this using Eq.
(9-1). However, we also need to know the servo’s starting voltage. The catalog
information in Figure 9-17 lists the starting voltage as 1.0 volt. For the Ks value in
Eq. (9-1), we substitute our earlier calculated value of 4 V/rad for KCx-ct There¬

fore,

(2)(1.0)
dead band -
(4X140.3)
= 0.0036 rad, or 0.21 deg

The figure of 0.21° means that the input control rotor on the CX has to be rotated a
minimum of 0.21° before causing movement in the servomotor. However, the
system’s hesitation to react to an input command doesn’t stop at the servomotor.
There is also the gear train’s backlash to contend with. Backlash is the dead band
of gear trains. It is the amount of rotation needed at the gear train’s input to just
cause rotation in its output. Backlash is caused by imprecise machining of gears,
and certain amounts cannot be prevented. Most likely, there is also slack in the
flexible mechanical coupling used in attaching the gear train’s output shaft to the
valve stem. All of these factors add to the overall “slop” within the control
system. This causes an increase in the system’s response time that is added to the
dead band of the synchro and servomotor combination. However, these addi¬
tional factors are difficult to obtain in data sheets and must be measured later after
the entire system is assembled. For the present, we neglect these other factors
beyond the servo’s dead band.
The complete system transfer function is:

40.78
KG (9-18)
5(1 + 0.0175)

Figure 9-20 is the open-loop Bode plot of our valve positioner system. Our phase
margin is within the recommended 40° to 60° range.
In the examples that we have covered so far, we have assumed that we could
easily obtain the components needed to make the system function the way we
wanted it to function. As any engineer or technician will verify, this is not a very
realistic situation. Often, the component you need the most is the component that
is the most difficult or impossible to obtain (no doubt a corollary of Murphy’s
Section 9-5 Example 4: An Oscillograph 289

w(Rad/Sec)
Figure 9-20 Bode plot for 40.78/5(1 + 0.0175).

law). The next example demonstrates how to make the best of a design situation
where only certain hardware components are available.

9-5 EXAMPLE 4: AN OSCILLOGRAPH

This example is somewhat similar to the one we discussed in Chapter 8 concerning


the voltage-reading chart recorder. Oscillographs are recording oscilloscopes.
That is, they produce a hard copy output of the waveform that is being analyzed.
Again, as in the previous design example, we deal with instrument-type servos in
our design because of the relatively light-duty applications involved.
In our design, we assume that we must work with an amplifier whose gain is
fixed at a value of 11. Because of cost restraints, we aren’t allowed to add any
additional amplifiers to the system. We assume, too, that the servomotor we have
to work with has a fairly high Kv. Its transfer function is:

/' ^out _ 35
Em s( 1 + 0.018s)

We also must work with a gear train having a gear ratio of 50: 1. Therefore, the
system’s overall transfer function that we have to work with is:
290 Chapter 9 Designing a Control System

1.1
KG (9-19)
^(l + 0.018s)

Figure 9-21 is the block diagram of our system.

In Out

Figure 9-21 Block diagram of oscillograph for Example 4.

Our first task is to look at the system’s Bode plot to see what it is we have to
work with. Because of the amplifier’s relatively low gain, we can, without looking
at the Bode plot, guess that we will probably be working with a fairly sluggish
system. Figure 9-22 shows us that our guess was correct. We see an extremely
sluggish system. The phase margin is greater than 80°. For the record, we can
calculate the damping ratio. Using the methods outlined earlier, we calculate a
factor of z = 1.343. Notice that this value exceeds 1. We really haven’t discussed
damping factors having values this large. We can guess, however, (and guess
correctly) that values this great do represent extremely sluggish systems. This

wCRad/Sec)

Figure 9-22 Bode plot for 7.7/5(1 + O.OI85).


Section 9-5 Example 4: An Oscillograph 291

extremely high value for z is undoubtedly due to the rather low Kx factor in our
transfer function. Since we can’t increase the gain of our system as we would
normally do (that was a stipulation given us in our problem), we instead have to
look at another alternative. We can install a phase-lag network. The phase-lag
network will increase the phase angle of the output (that is, make it more negative)
as compared to the system’s input, thereby decreasing the phase margin making
the system less sluggish. Looking back to Section 4-16.4 in Chapter 4, we see how
to construct this network (Figure 4-47). Following the general rule of thumb for
these networks, we make r2 = lOrj. Also, we can guess at a reasonable value for
the capacitor, C. Let’s make C = 1,000 pF. (For this high capacitance, we would
want to use a tantalum electrolytic capacitor to take advantage of its inherent
consistency with aging and good temperature stability. Bear in mind, though, that
we could have picked virtually any size capacitor for this application, since it
becomes a matter of developing the proper combinations of capacitance and resis¬
tances for creating the desired time constants. There are an infinite number of
these combinations.) We also pick a value of 100 ohms for R\. To determine the
value of R2, we set up the following equations:

Since t2 — 10Tl
and T\ = R2C
and t2 = (R i + R2)C
then t2 = 10Tl = 10^(1,000 x 10~6)
= (100 + /?2)( 1,000 x 10~6)

Therefore,

= 11.1 a

Since this is an odd-size resistor, we can approximate the value of R2 with a 10-
ohm resistor instead.
With the installation of our lag network, our transfer function becomes:

(1 + 0.0U)7.7
KG = (9-20)
5(1 + 0.1l5)(l + O.OI85)

Because of the complexity of this function, a straight-line Bode plot is not practi¬
cal. A programmable calculator that has been programmed for calculating the
gain and phase angle of our transfer function is recommended. The resultant
open-loop plot is seen in Figure 9-23. It’s interesting to note here that an approxi¬
mation to the actual Bode plot could have been made by merely cancelling the
(1 + O.OI5) expression with the (1 + O.OI85) expression, since they are approxi¬
mately equal. The simplified expression of KG = 7.7/[5(l + 0.115)] could then be
plotted. Figure 9-23 shows this approximation superimposed over the actual
curves. You can see that considerable inaccuracy is introduced at the high end of
292 Chapter 9 Designing a Control System

-45
-90 G
Ld
-135 9
-180 ~
_l
LD

Ld
C/0
<r
CL

1 10 100 1000 10000


(Rad/sec)
Figure 9-23 Bode plot for 7.7(1 + 0.01s)/.s(l + 0.11s)(l + 0.018s).

the gain curve, but near the zero gain crossover point there is a fair correlation
between the two curves. However, we are interested in the phase margin. We
see now that the phase margin has decreased to a respectable 54°. Since this is
within the desirable range, let’s now calculate our new damping ratio. Unfortu¬
nately, our transfer function is a cubic equation in its closed-loop form. However,
we can use the cancelled form, KG = 7.7/[s(l + 0.11s)], for a fair approximation
(see Example 8-5 in Section 8-6.4, Chapter 8). The damping ratio is found by
converting to the closed-loop form and using Eq. (8-5):

KG 7.7
1 + KG s( 1 + 0.11s) + 7.7
1
/0.11s2
s
' 7.7

(9-21)
/ s2 1.087s \
5 ' 8.3672 + 8.367 + '
z 0.543

This is an obvious improvement over the original damping ratio.


We have neglected the bandwidth up to this point in our design of the
oscillograph. Since we are concerned with the oscillograph’s ability to respond to
Section 9-6 Example 5: A Conditionally Stable System 293

a wide range of input frequencies, let’s see what sort of bandwidth we have with
the newly installed phase-lag network design. Using Eq. (8-17),

Mh = o>n Vl - 2(0.5432) + V2~-~4(0.5432) + 4(0.5434)


= 8.367 V 1 - 0.590 + VT- 1.179 + 0.348
= 8.367Vl - 0.590 + 1.081
= 8.367 Vl.491
= 10.22 rad/sec

This is the equivalent of 1.63 Hz. Judging from this very low bandwidth, our
oscillograph could only be used for the plotting of very slow cyclic waveforms,
approaching DC. Figure 9-24 is our revised block diagram showing the installed
phase-lag compensating network.

Figure 9-24 Block diagram of oscillograph with newly installed phase-lag net¬
work.

In summarizing our use of the phase-lag network, you have to be careful in


its application to a system. You can’t always depend on a phase-lag network, for
instance, to delay the system’s phase angle to make a sluggish system into a
desirably stable system. The system may be stable at the zero-gain crossover
point, but that same system may become unstable at some lower frequency. This
type of system is called a conditionally stable system. The next design example
demonstrates this problem.

9-6 EXAMPLE 5: A CONDITIONALLY STABLE SYSTEM

We now take our design example from Example 4 and modify it some¬
what. Instead of using an amplifier having a fixed gain of 11, we will change
that gain to, say, 300. This will give us a Kx of —, or 210. Our open-loop
transfer function now becomes,

210
KG = (9-22)
s{ 1 + 0.0185)

The Bode plot for this revised system is shown in Figure 9-25. The phase margin
of 32° indicates an unstable system as a result of the higher gain. We now add the
compensating network, (1 + 5)/(l + 105). Earlier, we mentioned that there are an
294 Chapter 9 Designing a Control System

PHASE ANGLE (DEG)


w(Rad/Sec)
Figure 9-25 Bode plot for the function 210/^(1 + 0.018s).

infinite combination of capacitances and resistances that could be used to satisfy


the time constant requirements for a compensating network. There are also an
infinite combination of t2 = IOtj to choose from. We picked the foregoing combi¬
nation merely to prove a point in this discussion. We want to see what happens
when we use this choice in combination with our transfer function, Eq. (9-22).
Adding the new phase-lag network makes our function become:

(1 + 5)210
KG (9-23)
5(1 + 105)(5 + 0.185)

Again, in order to plot this rather complex open-loop function, we want to use a
programmable calculator to make a point-by-point curve plot. Figure 9-26 shows
the results. At the zero gain crossover point, we find a phase margin of 68°,
indicating a somewhat sluggish system. But notice what happens if the gain were
to be reduced 43 dB. The phase margin becomes 35°, resulting in an unstable
system. We experience a situation that is somewhat similar to a spinning top. As
long as the rotational velocity is high the top remains stable; reduce its velocity
and it immediately becomes unstable. We have, as a result, a conditionally stable
system in the case of our top. In our control system, if the gain of a system is to be
varied for any reason during its operation and a compensating network is present,
be sure that you don’t have this type of existing situation just described. If you
do, you will want to avoid those particular gain settings. It’s interesting to note
Section 9-7 Example 6: A Digital Controller for a Hydraulic Press 295

wCRad/Sec)
Figure 9-26 Bode plot for the function 210(1 + ls)/s(l + 10s)(l + 0.018s).

that the inertially damped servomotor (the IDSM) has characteristics very similar
to the phase-lag network and is sometimes referred to as a phase-delaying de¬
vice. It is possible to have conditionally stable systems using phase-lead net¬
works. In any case, constructing the system’s Bode plot will uncover any unique
characteristics that may require special design or operating attention.

9-7 EXAMPLE 6: A DIGITAL CONTROLLER


FOR A HYDRAULIC PRESS

We haven’t talked about digital systems up to this point. They can become rather
complex, electronically. However, because of certain desirable characteristics
(which we discuss later), digital automatic control systems are becoming very
popular in industry. In this next example, we won’t delve into the electronic or
digital theory behind this particular system. It’s more important at this point to
simply understand the overall basics as to how the system operates. The digital
circuitry can be studied in other books (see the references at the end of this
chapter).
Figure 9-27 shows the schematic of an hydraulic press with an automatic
load control system. The load control is a keyboard and display where an opera¬
tor can type in the desired tonnage to be applied by the hydraulic press. The
296 Chapter 9 Designing a Control System

Tie Bars

Operates Only if 51 > B2


Shuts Down When B2 > By

Figure 9-27 Schematic of digitally controlled hydraulic press.

workpiece to be pressed is placed between the one movable plate attached to an


hydraulic ram or cylinder, and the lower stationary plate. When the hydraulic
pump is energized, the blind end (the end opposite the piston’s rod end) of the
cylinder becomes pressurized. This causes a clamping force to be applied to any
object placed between the two large plates. If the pressure becomes excessive for
any reason, we can open the regeneration valve. This will equalize the cylinder’s
pressure, thereby relieving the pressure between two large steel plates.
The press is held together, so to speak, by large circular bars or pillars of
steel, called tie bars. The larger presses usually have four of these arranged in a
square. The lower movable plate can then ride up and down on these bars.
We must come up with a method for determining the clamping pressure
created by the press. Assuming that the cylinder is pressurized, the movable plate
squeezes the object between its plates. While this is occurring, the clamping
cylinder exerts a force in the opposite direction by pushing against the top plate of
the press. You can envision this better if you picture yourself in place of the
cylinder, standing on the lower movable plate and pushing with your hands up
against the top plate. You must do this in order to create any squeezing action
Section 9-7 Example 6: A Digital Controller for a Hydraulic Press 297

between the two lower plates. The cylinder has to do the very same thing in this
instance. As a matter of fact, the tie bars forming the frame of the press will
become stretched during this squeezing action. As the cylinder expands between
the upper and lower plates, the tie bars experience a stretching action, which, if it
becomes excessive, could cause the bars to break or snap.
The amount of stretch or length change in any common metal, such as steel,
can be easily equated to the amount of force or pressure causing that stretch.
That relationship is:

(applied force) x (length of metal)


(9-24)
engt c ange (cross_sectional area) x (modulus)

The modulus, otherwise known as the modulus of elasticity, for steel is 3 x 107
lb./in.2 Therefore, if the applied force to a steel bar is known in pounds, the cross-
sectional area is known in square inches, and the length of the bar is known in
inches, then the bar’s change in length can be easily calculated using Eq. (9-24). It
now becomes a matter of determining how much the bars will stretch in our press
whenever the hydraulic cylinder is energized forcing the movable plate down onto
the bottom fixed plate.
The method we use for determining the amount of stretch taking place in the
tie bars is through an optical linear encoder (see Figure 9-28). This is a device that
can generate an electronic pulse for every 0.0001 in. of deflection of its plunger
rod. By counting these pulses we will know, within a tolerance of ± .0001 in., how
much movement of the plunger rod has taken place and, therefore, how much
deflection has taken place within each tie bar. This is assuming that the encoder is
attached to the press as shown in Figure 9-27.
The control portion of the system will work like this: With the output of the
optical encoder feeding into a frequency counter, we can count the encoder’s
output pulses. The count is then converted to a binary number, B2, and sent to a
digital comparator. This circuit is similar in principle to our differential ampli¬
fier. It has the ability to subtract one input from another and then present the
difference at its output. However, whereas the op amp is an analog device, the
digital comparator performs its subtracting digitally. And there is no amplification
involved either, as is generally the case with any digital circuit. The other input,
B\, to this comparator comes from a keyboard entry made by the press operator.
The output of the comparator circuit, which is in the form of a binary signal
resulting from the arithmetic difference between B\ and B2, is sent to a digital-to-
analog converter. The resulting analog signal is then amplified to drive a hydraulic
pump which, in turn, supplies hydraulic fluid to the press’s hydraulic ram. An
, important feature in the hydraulic circuitry of this system is the regenerative valve
shown attached to the side of the ram. This valve was mentioned earlier. When
the valve is closed, the cylinder operates normally. However, when the valve is
opened the pressure on the blind end of the piston is allowed to equalize the
pressure on the piston’s rod end; the downward force will then stop.
298 Chapter 9 Designing a Control System

Figure 9-28 The linear optical encoder. (Courtesy ofBEI Motion Systems Com¬
pany, San Marcos, Calif.)

We now need additional logic circuitry to operate the regenerative valve and
hydraulic pump. This circuitry will cause the regenerative valve to open only
when B2> By. Also, the hydraulic pump will not operate when B2> By, or it will
operate when By > B2. Logic statements such as these can become confusing at
first glance but are very common in digital circuits. Let’s study them a little more
closely to see how they make the circuit function. Let’s assume the operator
wishes to create a force of, say, 50 tons on a die that is being tested. The 50-ton
figure is entered on the keyboard and is automatically displayed on an alphanu¬
meric readout screen. At this same time, the decimal figure is converted to a
Section 9-8 Example 7: Bang-Bang Servo Systems 299

binary number (Bx) and sent to the digital comparator. Since at this time the press
is not operating, the pressure in the main cylinder is zero; therefore, B2 is zero.
Since Bx is greater than B2, the hydraulic pump turns on. As the pressure in¬
creases inside the cylinder, the magnitude of the binary number B2 also in¬
creases. Theoretically, when B2 equals Bx, the pump will shut down. However,
because of the existence of electrical and mechanical time constants within the
system, it is quite likely that the pressure within the cylinder will overshoot the
desired set point of 50 tons. If this happens, B2 will become greater than Bx and
the hydraulic pump will shut down. At this same time, the regenerative valve will
open, causing the pressure to equalize inside the cylinder. B2 will then become
less than Bx, causing the regenerative valve to open again and the pump to turn
back on starting the pressure building process all over again. With every cycle
that the system would make, the set point value of 50 tons would be approached
more closely until eventually the system will settle at that point.
Obviously, one drawback to this system design is that the system behaves
like one having a damping factor near zero. One step that could be taken to
reduce the number of oscillations (i.e., to increase the damping factor) is to place a
restriction in series with the regenerative valve to reduce the pressure reduction
rate inside the cylinder. Or, the speed of the hydraulic pump could be reduced to
reduce the pressure build-up rate within the cylinder. Another method, which
may be the most effective, is to build in a dead band area in the greater-than logic
of the regenerative value such that if B2 > Bx by, say, 2 tons, only then will the
regenerative valve open.
When dealing with digital circuits such as the one just discussed, the rules
for determining the system control characteristics change somewhat. This exam¬
ple was presented to introduce this point. There are methods for analyzing digital
systems that resemble the methods we have discussed for our analog systems, but
because of their complexity we don’t discuss them here. For one thing, determin¬
ing the transfer functions for large mechanical devices such as hydraulic compo¬
nents is a fairly complicated procedure. For cases like these, it is sometimes
quicker to build the system based on a preliminary design, and then through
experimentation determine the proper system components for the desired system
behavior. We discuss additional systems similar to this one in Chapter 10.

9-8 EXAMPLE 7: BANG-BANG SERVO SYSTEMS

Of the automatic control system types that presently exist, the bang-bang servo is
probably the simplest to design, operate, and understand. Because of this, this
- type of system is probably the most commonly used today for obtaining automatic
control. The bang-bang servo is a system that reacts to an input signal in a step
function manner. That is, the system is either fully on or fully off. There is
usually some sort of sensing device that monitors the output of the system. When
a particular level of output quantity is reached, usually determined by the system
300 Chapter 9 Designing a Control System

operator’s set point input, the system is turned off. As the output diminishes as a
result of being turned off, a minimum output quantity is reached causing the
system to turn back on. The cycle is then repeated over and over. Probably the
best example of a bang-bang servo is the thermostatically-controlled heating sys¬
tem for a home. Figure 9-29 is a diagram of just such a system. It shows how the

Set Point
(68° F)

>'

Thermostatic _ Heat
Furnace
Sensor Output

Heat Input

Feedback Path
(Room Air Distribution)
Figure 9-29 The thermostat.

furnace’s output is fed back to the thermostat that has been given a set point input
of 68°. In reality, we know the thermostat doesn’t actually sense the furnace’s
heat directly. If it did, the furnace would oscillate rapidly on and off, and as a
result the house would not become heated. Instead, the thermostat is usually
located away from a direct heater discharge and in an area where the room air has
had time to mix with the furnace’s output before actually becoming sensed. The
basis of operation of this system is the basis of all bang-bang servo systems. The
term servo is applied loosely in the case of our furnace example, since there are
really no servomotors present in the system.
Let’s take a closer look at the system just described and attempt to analyze
its behavior. Figure 9-30 shows a timing diagram of the furnace’s input and its
output as it arrives, so to speak, back at the thermostat. This then becomes the

Figure 9-30 Timing diagrams for a typical furnace and thermostat system.
Section 9-8 Example 7: Bang-Bang Servo Systems 301

furnace’s input once again. The first diagram, labeled System Output, shows the
step function characteristics of the furnace operation itself. The second diagram,
labeled System Input, shows the temperature fluctuations at the thermostat along
with the set point input (i.e., the thermostat setting). The first thing we should
note about these two diagrams is that they both have the same frequency. The
second important characteristic of these diagrams is the phase shift. Just as we
had a phase shift to contend with in our other servo systems, we have a phase shift
to contend with here. This out-of-phase relationship between the furnace’s output
and its input is shown as the phase angle 6 in Figure 9-30. This phase angle
amount is determined by several quantities:

1. the thermostat’s response time,


2. the thermostat’s hysteresis,
.
3 the house’s heat loss rate, and
.
4 the furnace’s ability to come up to operating temperature.

Perhaps as you read through the preceding list, you can see striking similarities
between this system and the others we discussed earlier. The frequency and
phase angle are obvious, but what about time constants? Can you see how we
could derive time constants for the house, thermostat, and furnace? Usually, you
don’t see time constant terms published for these items, but nevertheless, they do
exist and can be calculated if desired.
In item 2 of the preceding list, we mentioned the term hysteresis. Hystere¬
sis, in our application here, is the characteristic of a device or instrument that has
two different switching points for the same sensing quantity. These two points are
determined by whether the sensing quantity is increasing in magnitude or decreas¬
ing in magnitude. In the case of our thermostat, the thermostat may decide to turn
the furnace on at one temperature but may turn off the furnace at a different
temperature, even though the thermostat was given only one temperature set¬
ting. This discrepancy in switching temperatures is caused by such things as
static friction versus motion friction (two entirely different values), bearing clear¬
ances, direction of “play” in mechanical parts, and changes in physical dimen¬
sions due to temperature changes. These are just a few of the factors that figure
into the occurrence of hysteresis. Figure 9-30 shows where the hysteresis quan¬
tity occurs in our furnace-thermostat system.
In some cases, hysteresis is a desirable quantity. Take a look at Figure 9-30
again. If hysteresis were mostly eliminated from the thermostat’s design and
construction, how would this affect the cycling time of our furnace? Figure 9-31
' shows us the results. The room temperature would certainly be far more consis¬
tent, with very little fluctuation in magnitude. However, the price paid for this is a
furnace that cycles far more frequently and in shorter intervals. Most present-day
furnaces are not constructed to withstand that kind of high-duty rate or service.
As a result, many thermostats are now designed to allow the user to program in
302 Chapter 9 Designing a Control System

D
Q_
4->
D
o
CD
<J
CU
C
>- Time
D
LI-

Figure 9-31 Timing diagram for a furnace-thermostat system in which most of


the hysteresis has been removed from the thermostat.

just the right amount of hysteresis needed to produce the best heating effect for
the house in which the system is installed.
Another common application of the bang-bang servo system in use today is
the relay-operated motor positioning control shown in Figure 9-32. The relay coil
is a three-position relay. In the first position, the relay energizes the motor to

Figure 9-32 A relay-operated bang-bang servo system.


Section 9-8 Example 7: Bang-Bang Servo Systems 303

rotate in one direction, while the second position of the relay causes the motor to
rotate in the opposite direction. The third position, or middle position, is the off
position in which the motor is not energized. Notice that the motor is either fully
on in one direction or another or fully off, a characteristic of the bang-bang
servo. The advantage of the relay-operated system is that it is relatively inexpen¬
sive and that for a small power input to the system, one can control large amounts
of power at the output. The disadvantage with this system is the lack of accuracy
in positioning the motor. Since it is either completely on or completely off, there
is no modulating of the error or control signal. It is this finely tempered signal that
allows the motor to reach an exact position in the analog or digital type of servo
control.
The characteristics of the relay used in a bang-bang servo are very impor¬
tant. The off position in the three-position relay just discussed is the relay’s dead
band. Also, this relay will have a certain amount of hysteresis. That is, the volt¬
age needed to pull in its armature will be different from the voltage needed to
release the armature. This is typical of this kind of electromechanical device.
Figure 9-33 explains the characteristics of dead band and hysteresis and compares
them to a so-called ideal relay that has no dead band or hysteresis.

Output
(to Motor)

Forward
♦ +■

*>

Input ■

(to Relay Coil)

<*

Reverse

(a) Ideal (b) Dead Band (c) Hysteresis (d) Combined Dead Band
and Hysteresis

Figure 9-33 Relay characteristics used in a bang-bang servo system.

Now that we have discussed both the conventional (analog and digital) types
of automatic control system and the bang-bang servo automatic control system,
we should compare their output characteristics. A characteristic of the bang-bang
servo is its habit of excessive hunting as compared to the conventional systems.
Figure 9-34 compares the two general systems. Notice that in addition to the
excessive hunting of the bang-bang system, the final settling point is somewhere
within the dead band of the switching relay. The greater the dead band of a
switching relay, the greater the inaccuracy of the system’s positioning ability.
304 Chapter 9 Designing a Control System

Figure 9-34 Comparing response curves of a bang-bang servo system versus a


conventional control system.

There is also dead band within the conventional system, but not to the degree that
you have with the bang-bang system.
One way to reduce the amount of hunting and the frequency of oscillation in
a bang-bang servo is to introduce a certain amount of viscous damping to the
motor. Unfortunately, when this is done the settling time is increased. The over¬
all speed of the system is slowed.

SUMMARY

The designing of an automatic control system is heavily dependent on good engi¬


neering catalog data obtained from the components’ manufacturer. Without this
information, the design work can become very frustrating. Designing these sys¬
tems not only requires a good knowledge of automatic control theory; it also
requires a fair knowledge of physics. Without physics, designing a control system
to operate and control a mechanical device that has to obey certain physical laws
becomes an impossibility.
Digital controllers are becoming more and more popular as automatic con¬
trolling devices. The reliability of digital components along with their miniaturiza¬
tion is making them a very attractive alternative to the analog type of control
system. Example 6 in this chapter was a good representation of a typical digital
control system.
Bang-bang servo systems are yet another alternative to the analog and digital
control systems. They have the advantage of being simple in design and inexpen¬
sive to manufacture. Their greatest disadvantage is their inaccuracy of operation
as compared to the other systems.
References 305

EXERCISES

Because of the complexity of automatic control design problems, the following is


a list of suggested design problems. You will need to obtain catalog data from the
various servo components manufacturers in order to undertake these suggested
projects. The final submitted design should show all data calculations, block dia¬
grams, curve plots, and design catalog data from which all component perfor¬
mance data was obtained.
9-1. Design an x-y curve plotter that can be used with a computer for plotting
graphs and curves.
9-2. Design a positioning system for locating a tool on an NC (numerical control)
machine.
9-3. Design a speed control system for controlling the rotational velocity of a
rotating tool for an NC machine.
9-4. Design a cruise control system for an automobile taking into account vari¬
able factors such as road grade, windage, and in-town driving.
9-5. Design a flow control system for a beverage bottling company’s filling
system.
9-6. Design a radar tracking system that locks onto a high-altitude flying object
such as an airplane.
9-7. Design a positioning system that controls the rotating or swivelling of a
robot’s main body or torso as it reports to the various positions for perform¬
ing manufacturing functions.
9-8. Design a bi-level temperature controller (i.e., one that controls both heating
and cooling) for an environmental chamber that must have its internal tem¬
perature maintained at a specific value. Temperature range requirements
are between 0°F and 200°F.
9-9. Design a remote positional controller for a robot arm located in a hazardous
environment and used while the arm’s operator is in a safe location several
miles away.
9-10. Design a camera positioner that can lock onto a faint light source at night for
the purpose of tracking and recording the movements of that light source.

REFERENCES

Baeck, Henry S., Practical Servomechanism Design, New York, N.Y.: Mc¬
Graw-Hill, Inc., 1968.
Deem, Bill R., Kenneth Muchow and Anthony Zeppa, Digital Computer
Circuits and Concepts, Reston, Va.: Reston Publishing Co., 1980.
306 Chapter 9 Designing a Control System

Dorf, Richard C., Modern Control Systems, Reading, Mass.: Addison-Wesley


Publishing Co., 1980.
Malvino, Albert Paul and Donald P. Leach, Digital Principles and Appli¬
cations, New York, N.Y.: McGraw-Hill, Inc., 1975.
Sandige, Richard S., Digital Concepts Using Standard Integrated Circuits,
New York, N.Y.: McGraw-Hill, Inc., 1978.
Process Control
Systems

10-1 INTRODUCTION

Chapter 9 showed us how to go about designing some fairly basic automatic


control circuits. These designs were just complicated enough to give you a flavor
as to what designing is all about. In this chapter, we discuss how various process
control systems work. A process control system is comprised usually of one or
more automatic control systems dedicated to regulating the manufacturing of a
particular product. Enough variety of processes are presented here to give you an
idea of how other automatic control engineers and technicians apply the principles
discussed in earlier chapters of this book. In many cases, the application of
automatic control theory to making a particular process work is not obvious. You
have to remove machine shrouds, open control box doors, and dismantle parts in
general to see how the controls theory is applied by the processes’ system de¬
signer. While we can’t do that here, the next best thing is to describe as earnestly
as possible how the systems work and how the principles behind them operate.

10-2 ROBOTIC CONTROL SYSTEM

, Figure 10-1 shows the type of robot that we discuss here. This is a jointed-arm
spherical robot. It is used in the automotive industry for welding car frames on an
assembly line. Because of its ability to reach into tight areas and corners, this
type of robot is used to reach those areas within the frame that are difficult for
human welders to reach. This robotic design has what is commonly referred to as

307
308 Chapter 10 Process Control Systems

Figure 10-1 Jointed-arm spherical robot. (Courtesy of Cincinnati Milacron, Cin¬


cinnati, Ohio.)

six degrees of freedom (Figure 10-2). With these freedoms of movement, this
particular robotic design has access virtually to every point or location within a
certain sphere or volume surrounding the robot. But how does the robot know
how far out it is supposed to reach or how far should it rotate, and how does it
know when it is finally there?
Basically, the control mechanisms for all six motions within the robot could
be designed to be roughly the same. We look at one method for controlling the
positioning of the robot’s arm. Figure 10-3 shows the mechanism for controlling
the amount of swiveling of the robot’s waist or base; this amount is the 6 indicated
in Figure 10-2. This system is similar to the hydraulic digital controller example
described in Section 9-7. The operator merely keys in the coordinates for the
robot to swivel to. This information may also be supplied to the robot by means of
a computer program or software associated with an attached computer or pro¬
grammable controller. (We talk about programmable controllers in Chapter 12.)
Section 10-2 Robotic Control System 309

Figure 10-2 The six degrees of free¬


dom for a robot.

Figure 10-3 Diagram for a robotic keyboard input control system.


310 Chapter 10 Process Control Systems

The digital comparator then calculates the binary arithmetic difference (the error)
between the desired position (the set point) and the swivel’s actual location as
read by the optical encoder attached to the swivel’s drive shaft. The magnitude
and sign of this arithmetic difference determines the direction of rotation of the
swivel motor.
In order for this system to function properly, there has to be a preselected
“home” position for the robot so that the encoder doesn’t put out ambiguous or
conflicting information.
Let’s look at an example. Using Figure 10-2, imagine taking the robot’s arm
and rotating the arm 180° so that it extends in the opposite direction. This would
place the robot’s arm to the rear. Let’s assume that this is our homing position
and it is facing, say, north. At this location we adjust our encoder so that the
frequency (pulse) counter is outputting a zero count. That is, the binary number
output is 00000000000000, or fourteen zeros. (The number of zeros indicates in
this case that the maximum binary number the counter can read is
11111111111111, or fourteen ones. The decimal equivalent of this binary number
is 214, or 16,384.) Now, let’s swing the robot’s arm in the opposite direction from
the home point and swing it a full 360° until we are once again at home but arrived
there from the opposite direction. Let’s say that our frequency counter now
shows a binary reading of:
10010011010001

The decimal equivalent of this number is 9,425. An encoder was selected to


generate this number by choosing one having an encoder disk 3 inches in diameter
and having 1,000 lines per inch. That is, this encoder will generate 1,000 pulses
for every inch of movement around its circumference. When installed on the
robot, it can be swiveled, in theory, to within ±0.001 inch of the desired location.
Some sort of mechanical stop would have to be installed at home so that the
robot’s rotation can’t go beyond this point, thereby causing the encoder reading to
“roll over” to zero. If this happened, the robot’s comparator output would be¬
come confused, causing uncontrollable rotation.
Let’s make a further assumption that a positive error signal coming from the
digital comparator will drive the swivel motor clockwise, while a negative error
signal will cause the motor to drive the robot’s base counterclockwise. Having
established the full range of swiveling motion and binary output numbering of our
encoder, let’s swing the robot to some intermediate location to where the encoder
output, according to the frequency counter, is now reading binary
01010011010001. If a binary number of, say 00001011011000 is entered into Figure
10-3’s comparator by means of software or by keyboard, the arithmetic difference
between this number and the binary number representing the robot’s location
becomes:
01010011010001
- 00001011011000
01000111111001
Section 10-2 Robotic Control System 311

Since the remainder is positive, the swivel motor will rotate clockwise and will
continue to rotate clockwise until the remainder reduces to zero. In reality, be¬
cause of the inertia of the robot’s body, what is most likely to happen is that the
motor will overshoot the zero-difference mark somewhat, therefore causing a
negative error difference to be generated. The motor will then reverse itself in an
attempt to reduce the error to zero once again, but again overshooting the zero
mark somewhat. The error, however, should be smaller. The motor’s polarity
will once again become reversed, therefore starting the homing or hunting all over
again, each time coming closer to the zero mark. This is similar to what takes
place in an analog control system. However, because of the definite on or off
state of each pulse being counted, there is a possibility that the motor could
continue correcting itself either side of zero, oscillating back and forth perhaps as
little as ±0.001 inch. In other words, because of the very fine resolution of the
optical encoder, and because of the possibility of the robot’s swivel movement
landing very near the edge of one of the optical lines on the encoder, the encoder
could easily become confused as to whether or not to generate a pulse. It there¬
fore oscillates between the on state and the off state indefinitely or until it receives
a new command.
To prevent the hunting action just described from taking place, an artificial
dead band can be created. If the resolution of 1,000 pulses per inch isn’t needed,
merely going to a lower resolution on the encoder will lessen the problem. If the
resolution can’t be sacrificed, viscous damping can be introduced to the swivel
base. One method is to use a hydraulic pump that the swivel must continuously
work against. Unfortunately, this is a waste of energy, but at least it is an effec¬
tive cure without sacrificing accuracy. Another method is electronically to intro¬
duce a dead band. This is done by lifting the restrictions somewhat as to what
constitutes a zero at the comparator’s output. The comparator can be duped into
thinking that if it gets to within binary ±00000000000011 (±3 pulses, or 0.003 in.),
the error is as good as zero. Obviously, the price you pay for this artificial dead
band is accuracy. Nevertheless, this is another effective way to prevent hunting
at the zero error point.
Because of the similarities between the digital and analog forms of automatic
control, let’s take a look at the analog form to see if there are any advantages or
disadvantages. The analog form by its very nature does not deal with discrete
numbers or levels. In other words, there is no such thing as being close to the
edge of an adjacent value or number as in a digital system. The voltage values are
continuous. Consequently, there is no digital hunting due to the lack of dead band
as in the case of very high-resolution digital systems.
Digital systems do have significant advantages over the analog systems. It is
relatively easy to filter out extraneous electrical noise from digital systems. For
the most part, electrical noise is analog in nature; therefore, building filters to
separate this interference from the desired digital signals is relatively easy. Trying
to separate analog noise from desired analog signals, on the other hand, can be
difficult and frustrating.
312 Chapter 10 Process Control Systems

Another advantage of digital control over analog control is that digital con¬
trols interface naturally to the computer. Since the computer is a digital device, it
uses the same language as the controlling circuitry, therefore eliminating any need
for analog-to-digital or digital-to-analog conversions.

10-2.1 Problems with Using Digital Encoding

With the use of digital encoding, such as the digital optical encoder, new
problems are created concerning generating the output signals. In the aforemen¬
tioned robotic control, we assumed that the swiveling of the robot’s waist was an
uninterrupted movement, the movement being registered by the number of pulses
generated by the attached optical encoder. What happens if the robot accidentally
strikes an object causing it to stop momentarily, reverse from the rebound, and
then start up again in the original direction? The momentary rebounding or re¬
versing adds additional pulses to the output count of the encoder. The robot could
easily become lost as a result. It will have lost track of its command by reporting
to the set point too early. This is where absolute encoding is used. Absolute
encoding has the capability of keeping track of any reversals that may take place
by subtracting any attempts to add additional counts due to these reversals. As a
result, the robot can keep track of its location and know absolutely where its
position is.
Another form of absolute encoding is where the encoder outputs a binary
number instead of a single pulse. The binary number itself acts as an address,
therefore telling the robot exactly what position it is in. However, there are
problems with this system too. Look at Figure 10-4. Here we see a simple sche¬
matic of a typical encoder of the binary number-producing type. Looking at the
left-hand illustration, notice what happens as the encoder wheel rotates counter¬
clockwise. Theoretically, all three photocells should switch to whatever the new
binary number is. According to the illustration, photocell 1 should remain low,
photocell 2 should switch high, and photocell 3 should switch low, forming the
binary number 010, or decimal 2. In reality, however, the three photocells will not
switch to their new states simultaneously. There will be intermediate or spurious
values of binary numbers generated as the division line on the encoder separating
the two binary numbers 001 and 010 passes underneath the photocells. Other
possible random numbers may be 000 and OIL
To get around this problem, a different sort of counting code is used. It’s
called the gray code. The gray code is so arranged that when counting up or down
from one decimal equivalent to the next, there is only one binary digit or bit that
changes to form that new number. Figure 10-5 shows a simple encoder wheel
layout using the gray code. As a result, as the wheel is spun beneath the sensors,
the only possible random numbers that could be generated during the numeral
change are equal either to the present number or the number that is just about to
be generated. This is a very simple scheme and a very effective one. There are
Section 10-2 Robotic Control System 313

2 3
r\
Light Sources

i l l
_1_1_1_
1 1

u 1 i
i
l
l

Photo Sensors

Top V iew

Three-bit 111 000

Light Source
and Sensor
Locations

Figure 10-4 A three-bit binary encoder


system.

100 000

Figure 10-5 A three-bit Gray code


wheel.
314 Chapter 10 Process Control Systems

other methods used in addition to using the gray code, but this one is probably the
most popular at the present. Whatever method is chosen, however, will require a
means for converting the encoder output to a binary number to remain compatible
with the rest of the operating system.
Virtually all robots used in industry are part of some larger automatic control
system. The robots themselves, of course, are comprised of many automatic
control systems working in concert to create the proper movements within the
robot. There are basically three types of power sources that a robotic designer
has to work with: hydraulic, pneumatic, and electric power sources. Each of
these sources has its own unique hardware design problems as far as designing the
needed control systems. Convincing arguments can be made for using any of
these systems, and consequently, you will find all three sources used extensively
in industry.

10-3 VISION SYSTEMS

Using vision in automatic control systems is perhaps one of the most exciting
aspects of controls design today. It is a field that is developing rapidly and
presents many technological challenges to the engineer. A vision system is a
system that uses an image of an object to extract data and to cause another system
or other systems to react. Stated more simply, a vision system is like a TV system
that has been programmed to do some sort of function when it sees something it
recognizes. Much of the vision system development has been performed by the
military for military applications. Specifically, low-flying missiles have been
given programmable vision systems to recognize certain flight paths and target
areas. What the camera sees determines the reaction of the automatic control
system controlling the missile’s flight direction.
To understand how vision control systems operate, we must first understand
how the vision or camera’s optical system works. The best way to approach this
topic is to first understand how a photographic camera works. Figure 10-6 shows
a simple schematic of a still photographic camera showing the lens system that
projects the image of an object onto the camera’s film. In place of the film,
however, let’s place a specially made-up plate comprised of many tiny photo-
• cells. Each photocell will be either turned on or turned off depending on the
object’s light intensity at the location of that particular cell. If the cell is on, we
can call its particular signal a 1. If the cell is off, that is, there is little or no light
falling on the cell, the cell’s output is a 0. (Remember, the 1 can represent any
voltage level, not necessarily 1 volt, while the 0 represents some lower voltage,
not necessarily 0 volts.) We then scan each row of the photocells using a special
electronic circuit that interrogates each cell to determine which state or level the
photocell is in. In other words, this scanning circuit will find out if the photocell is
high (1) or low (0). We can then represent our object, in this case the letter A, as a
Section 10-3 Vision Systems 315

Film Plane

Camera
Body

Lens System

Figure 10-6 How a still camera forms an image.

series or mosaic of Is and Os arranged as in Figure 10-7. The resultant image is


crude but nevertheless recognizable. Obviously, better resolution can be obtained
by using more photocells. What is interesting about our vision system is that
since we have digitized the image into Is and Os, much like the digitizing of
computer data, we can easily store this data onto magnetic tape or disks just as we
do with computer data.
Now that we have a method of creating digitized images, we must look at a
method for using this information in an automatic control system. Perhaps one of
the most active areas of vision development right now in industry is the field of
parts recognition. Figure 10-8 shows a typical example. Here we see a system

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 1 1 1 1 0 0 0 0 0 0
0 0 0 0 0 1 1 1 1 1 1 0 0 0 0 0
0 0 0 0 0 1 1 0 0 1 1 0 0 0 0 0
0 0 0 0 1 1 0 0 0 0 1 1 0 0 0 0
0 0 0 0 1 1 0 0 0 0 1 1 0 0 0 0
0 0 0 1 1 1 1 1 1 1 1 1 1 0 0 0
0 0 1 1 1 1 1 1 1 1 1 1 1 1 0 0
0 0 1 1 0 0 0 0 0 0 0 0 1 1 0 0
0 0 1 1 0 0 0 0 0 0 0 0 1 1 0 0
0000000000000000
0000000000000000
0000000000000000
0000000000000000 Figure 10-7 Digitizing an image. A bit
0000000000000000 map for the letter A.
316 Chapter 10 Process Control Systems

Figure 10-8 Automatic control system using vision system for parts recognition.

that has a capability to compare parts flowing past a checkpoint on an assembly.


An overhead video camera takes a snapshot of each part, and the recorded image
is sent to a processing center. At this location, the camera’s image is compared to
a prerecorded “standard” image of an identical part for a bit-by-bit comparison.
If the two images compare, a signal is automatically sent to the robot telling the
robot to let that particular part pass by. If, during the comparison process, the
two images don’t match, a signal is sent to the robot telling it to reach into the
assembly line flow to pick out that part. The robot then places it in a holding area
for further inspection.
In order to better understand the system, let’s assume the following prob¬
lem: Let’s suppose that we are trying to separate As from Bs on an assembly line
in a company that manufactures letter cutouts for an advertising firm. Again, we
have a robot that will reach over to pluck out the As from the Bs when com¬
manded to do so. Actually, our robot can be a simple pick-and-place type ma¬
chine rather than the much more expensive articulated-arm type depicted in Fig¬
ure 10-8. In any event, each letter is scanned by the camera and the image
information is sent to a signal processing circuit. This circuit then converts the
camera’s electrical data into binary data and prepares the data for a direct match¬
up with the standard binary information stored in the computer. This comparison
is done on a line-by-line, bit-by-bit basis, as mentioned earlier. If any of the bits
don’t match, as in the case of comparing an A to a B, or if the A is out of tolerance
as a result of a portion or all of it being either too large or too small, an error signal
Section 10-4 Camera Leveler System 317

is sent to the robot to reject the part. Looking at Figure 10-9, which shows the
letter B and its bit map, it’s fairly obvious that the third row of bits down from the
top of the bit map doesn’t match with the third row in Figure 10-7. The binary
number is entirely different in value.

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 1 1 1 1 1 1 0 0 0 0 0 0 0
0 0 0 1 1 1 1 1 1 1 0 0 0 0 0 0
0 0 0 1 1 0 0 0 1 1 0 0 0 0 0 0
0 0 0 1 1 1 1 1 1 0 0 0 0 0 0 0
0 0 0 1 1 1 1 1 0 0 0 0 0 0 0 0
0 0 0 1 1 1 1 1 1 0 0 0 0 0 0 0
0 0 0 1 1 0 0 0 1 1 0 0 0 0 0 0
0 0 0 1 1 0 0 0 1 1 0 0 0 0 0 0
0 0 0 1 1 1 1 1 1 1 0 0 0 0 0 0
0 0 0 1 1 1 1 1 1 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 Figure 10-9 Bit map for the letter B.

In order for a proper image comparison to be made, it’s necessary that the
camera be placed directly overhead of the viewed object. This is so that there will
be no visual distortion of the object’s dimensions. Quite often, only silhouettes of
the viewing object are used, since the outline represents the overall dimensional
shape. The present technology is limited as far as being able to recognize details
inside the object’s profile lines; however, it will be just a matter of time before that
will change.
The individual bit representing the light’s intensity at a particular location in
an image is called a pixel. Many times, it takes thousands of these pixels to
compose just one image, depending on the image’s physical size. It’s possible,
however, to use a completely analog system for parts recognition purposes. This
means using ordinary video cameras. In some cases, they can be used just as
effectively in an automatic control system as can the digital systems. It is obvi¬
ous, though, that somewhere along the system’s design layout it will be necessary
to convert this information to digital information for computer analysis and
storage.

10-4 CAMERA LEVELER SYSTEM

This next system we discuss is interesting because of its unusual design. In the
entertainment field of movie-making and television, there is a need to maintain the
318 Chapter 10 Process Control Systems

levelness of the recording camera that has been mounted on a moving vehicle.
Obviously, unless done for special effects, you don’t want the camera to record
any of the dips or rises that take place on not-so-smooth-or-level roads. Also, in
the case of hand-held cameras, or cameras pushed on dollies, you want to be able
to maintain levelness with the horizon at all times while shooting a scene.
One of the earlier solutions to this problem was strictly mechanical in its
application. The solution involved using a gyroscopic device that spun rapidly to
create the necessary opposing forces for maintaining a horizontal platform. How¬
ever, considerable mass was necessary to produce the opposing forces, and con¬
sequently, the entire system was massive and cumbersome. In addition, some
sort of on board power supply was needed to keep the gyroscope spinning.
However, these systems were surprisingly effective, and some are probably still
in use today.
A system that will enable us to maintain camera levelness is shown schemat¬
ically in Figure 10-10. We see a camera-mounting plate that is pivoted in its

Figure 10-10 Camera leveling system.


Section 10-4 Camera Leveler System 319

center. The plate can rock in any direction about this pivot. Two servomotors
are attached to the plate at the locations shown in Figure 10-10. Either or both
motors, when energized, will cause the plate to tip. The plate’s levelness is deter¬
mined by two plumb sensing devices attached to the plate at right angles to each
other. These devices are essentially nothing more than potentiometers whose
wipers have become plumbs. When the plate is absolutely level, each pot outputs
0 volts to its respective servomotor. As the pots are tilted according to the tilt of
the mounting plate, each pot puts out a proportionally varying DC voltage to its
servo. This up-and-down tilting, by the way, is referred to as the object’s pitch.
An object’s roll is the rolling over of the object; this is controlled by the pot which
is at right angles to the first pot. Tilting in one direction causes either pot to output
a positive DC voltage, while the opposite direction produces a negative DC volt¬
age. The resulting actions of the two pots and their respective servomotors cause
the mounting plate to remain level.
A system like this must be able to react quickly to sudden changes in its
orientation. The system’s components must have small time constants yet be able
to move with virtually no hunting. The viewing audience watching the film or
video tape will, otherwise, likely detect the slightest movement of any scene
pitched relative to the horizon.
Up to this point, we have discussed a leveling system to accommodate a
moving platform’s orientation with respect to the horizon. But what about side-
to-side motion relative to a fixed point out on that horizon? Is there any way to
compensate for this side-to-side motion? One method involves using another
potentiometer having a heavily weighted wiper that lies flat on the platform plate.
Because of the massiveness of this wiper, its inertia will resist any sudden change
in a sideways direction. Essentially, what happens is that the resistance windings
of the pot will move beneath the wiper, which remains pointed straight ahead.
In order to accommodate both the up and down servomotor action for
proper pitch and the sideways motion (this motion is often referred to as the yaw
of a moving object), a modification to our camera platform has to be made. A
second plate must be added that is mounted over the first plate and is free to rotate
relative to the first plate. The sideways servomechanism is fixed to the first plate
and pushes and pulls on the second plate to maintain its sideways orientation, as
shown in Figure 10-11. The camera equipment is mounted on this second plate.
So far, we have discussed an all-electronic system using electrical servomo¬
tors. It is also possible to use a system consisting of either hydraulic or pneumatic
actuators in place of the servomotors and their associated rack and pinion drives.
One unique design using pneumatics involves the using of air bags mounted in
place of the two servomotors under the bottom-most plate. However, for better
reaction time and control, a third and fourth air bag are mounted opposite the first
two. As one bag is inflating the opposite bag is deflating, and vice versa. How¬
ever, in order to get this system to function properly, we have to have some sort of
valve switching arrangement to shuttle air back and forth between the opposing
bags. This is where the proportional-type servo-valve can be used. This is a
320 Chapter 10 Process Control Systems

Mounting Plate

Figure 10-11 A system for controlling yaw. The mounting plate shown is
mounted on top of the plate shown in Figure 10-10.

valve that is electronically controlled. It contains an internal shuttle or poppet


whose position is precisely controlled by electronic pulses supplied from another
circuit. The position of this shuttle determines the amount of air that is allowed to
pass through the valve. The circuit producing the electronic pulses is told how
many pulses to produce by the output of the attitude-sensing elements used earlier
in our leveling system. The second plate that controls the camera’s yaw can be
controlled similarly. An added bonus of this type of system is that by using air
bags, you obtain a high degree of shock absorbing capability that is not available
with any of the other systems.
In place of the pneumatic air bags, hydraulic or pneumatic cylinders may be
used. An advantage to using these types of actuators, especially the hydraulic
actuators, is that tremendously heavy loads may be positioned using them. It is
difficult, however, to obtain the high degree of positioning accuracy and control
that you have with the servomotor. This is true in many automatic control sys¬
tems. However, pneumatic and hydraulic systems tend to be somewhat less ex¬
pensive, and with the advent of recent precision servo-positioning valves, hydrau¬
lic systems are becoming performance competitive. Keep in mind, though, that
with any hydraulic or pneumatic system, you have to have a pump or compressor
on board the system. This increases the noise, cost, and weight of the system.
Section 10-5 Paper Processing System 321

10-5 PAPER PROCESSING SYSTEM

Like many manufacturing processes, the paper industry is heavily dependent on


automatic feedback control systems. Automation has not only allowed the de¬
crease in manufacturing tolerances on many products (resulting in higher quality),
but automation has also removed operating personnel from undesirable or hazard¬
ous work situations which once required manual operation. Let’s take a look at a
typical paper manufacturing setup where we see the wet, newly formed paper
entering its final stages of production. This particular manufacturing process is
interesting because, unlike the others we have discussed, the response times (and
therefore, time constants) are quite long. This is characteristic of any mechanical
system using very large, massive system members for its operation. Because of
this, considerable time may pass between the time the error signal is generated
and the time a response is noted in the process. We may be looking at times
measured in minutes or even portions of an hour. Nevertheless, we will still
experience damping factors in the 0.4 to 0.7 range, hopefully, as we would like to
have them. It’s just that we will have to wait that much longer for our transient
curve to develop. As a matter of fact, some system designers, when given the
option of selecting a damping factor, will opt for values greater than 0.7 in the case
of very slow systems such as this one, to reduce over-shoot from producing a high
out-of-tolerance product in addition to a low out-of-tolerance product. (This prac¬
tice is considered questionable by some system designers.) In reality, whenever a
response is generated by a sensor, as we discuss shortly, the error signal is really
not a step input. It is actually a gradually increasing or gradually decreasing error
signal and is really considered to be a ramp-type input. This alters our overall
transient curve form appearances from the ones we have been studying, but our
overall system behavior is much the same. The study of ramp inputs is covered in
advanced automatic controls texts. (Note: a ramp function is nothing more than
the mathematical integral of a step function. And if you had a parabolic-type
function for an input to your system, the parabolic function is merely the integral
of a ramp-type input.)
Figure 10-12 shows the calender roller operation where the paper’s thick¬
ness, moisture content, and weight quantities are all controlled automatically.
Calender rollers are used for controlling paper thickness and humidity in the final
manufacturing stages of sheet paper. In some instances, the final paper product is
rolled onto a final take-up roller for storage before finally being shipped to the
customer. What we are interested in are the feedback control systems used for
controlling the paper’s various aforementioned quantities.
Let’s look at the first quantity, paper thickness. To control paper thickness,
;we must first have a transducer that can measure paper thickness while the paper
sheet is moving rapidly from roller to roller. A roller caliper would do just that for
us. This is a special caliper (Figure 10-13) having a set of rollers attached to each
caliper end with each end positioned on either side of the moving sheet. The
caliper rollers are adjusted so that the top roller rests on the bottom roller through
322 Chapter 10 Process Control Systems

Figure 10-12 The calender control of a paper rolling operation.

the sheet. A spring tensioning mechanism is used to maintain proper press be¬
tween the rollers at all times. The amount of caliper displacement due to the
sheet’s thickness is detected by a displacement sensor, consisting of a capacitive
sensor, linear variable differential transformer, or optical encoder. The thickness
is then automatically recorded and compared to the desired set point thickness.
The calender roller pressure is then adjusted by varying the oil pressure inside the
rollers. The rollers are actually flexible to allow their diameters to be changed.
Proper humidity must be maintained during the paper manufacturing process
in order to maintain proper roller feed rates in the calender system. A too-wet
condition will cause the paper sheet to become weak and to tear. A too-dry
condition will cause static electricity to develop, setting up a potential for possible
fire. Also, the curing of the paper sheet must be done at a specified rate to

Figure 10-13 Roller caliper for measuring and monitoring sheet paper thickness
during rolling operation.
Section 10-5 Paper Processing System 323

maintain a certain fiber strength within the paper’s material. Humidity affects this
rate. Therefore, humidity sensors must be placed at strategic locations through¬
out the paper rolling process to monitor the relative humidity. Two types are
used: One type consists of a horse-hair or human-hair hygrometer where the
expansion and contraction of a hair strand is a measure of relative humidity
(Figure 10-14). The other type is a chemical process where either the electrical
resistance of a hygroscopic material (a material, such as calcium chloride, that
absorbs water) is measured to determine the relative humidity, or there is a
current flow generated due to a chemical ion exchange process (Figure 10-15), or a
dielectric constant varies proportionally with humidity. The signal output infor¬
mation from the humidity sensors is sent by telemetry back to the paper driers,
tension rollers, calender, and to the plant’s environmental air conditioning sys¬
tems for the proper humidity adjustments. In the case of the calender system, the
calender rollers’ temperature can be regulated by controlling the heating or cool¬
ing of oil inside the rollers. This, in turn, controls the relative humidity or mois¬
ture content of the paper sheet. It’s obvious, at least in this particular control
system, why response times could be considered slow compared to what we have
been studying. Heating or cooling the volume of oil used in the calender could
produce some rather large time constants.
The weighing of paper while still in motion coming off of the calender rollers
can be an interesting measuring and control problem. One method is quite simple
but disruptive to the operation of the manufacturing system. The system is simply
shut down momentarily while sheet samples are cut from the paper stream and
weighed. A less disruptive system has the take-up roller weighed as it revolves.
It is necessary to know exactly how much length of paper has been taken up on
the roll, but that figure is not difficult to obtain with the proper metering devices.
What is difficult, though, is the adjusting of any needed paper weight. The re¬
sponse time for this operation is extremely long, and because of this, this particu¬
lar operation is done manually rather than automatically. While the weighing
operation is monitored regularly, the weight adjusting is usually done manually by

Pivot

Figure 10-14(a) Schematic of an hair


hygrometer used for measuring relative
humidity. (Ronald P. Hunter, AUTO¬
MATIC PROCESS CONTROL SYS¬
TEMS: Concepts & Hardware, 2nd.
Ed., copyright 1987, p. 103. Reprinted
by permission of Prentice Hall, Inc.,
Englewood Cliffs, N.J.)
324 Chapter 10 Process Control Systems

(b)

Figure 10-14(b) Commercially constructed hygrometer. (Photograph courtesy


of The Foxboro Company, Foxboro, Mass.)

adjusting the pulp mix back-up at the front of the manufacturing operation. This is
assuming, of course, that the sheet is of the correct thickness. Otherwise, the
weight variance could have been the result of a change in the paper’s thickness.
There are areas in this manufacturing process that demand fast response
times for the prevention of system failure. For instance, fast response times are
needed to compensate for changes in sheet tension as the sheet flows through the
Section 10-5 Paper Processing System 325

Figure 10-15 A sulfonated polystyrene ion-exchange sensor for measuring rela¬


tive humidity. (Courtesy of General Eastern Instruments Corp., Watertown,
Mass.)

calender rollers. If the tension is too great, there is a risk of sheet breakage. If the
tension is not great enough, the take-up roller will become too loosely wound with
the paper product. Sheet tension is adjusted and maintained by electrically or
hydraulically driven tension rollers shown in Figure 10-12. A tension sensor in the
form of a broad pressure-sensing roller that rides on the moving sheet detects any
change in sheet tension. Any change is instantly translated to a compensating up
or down movement of the tension rollers. The system behavior of this particular
automatic control system closely resembles the classic systems that we discussed
in earlier chapters.
Another important automatic control system has to do with variable roller
speeds. As the paper sheet comes off of the calender onto the take-up roll, the
take-up roller must be able to modulate its rotational velocity to match that of the
sheet. Assuming that the linear velocity of the sheet remains constant coming
from the calender, the take-up roller must be able to reduce its velocity propor¬
tionally as the roll becomes larger. This can be done with a sensor that senses the
roll’s increasing diameter. This can be a roller-caliper that rides on top of the
, take-up roll and whose displacement from the center of the roll is detected by a
linear encoder. As this linear displacement increases as the roller becomes larger,
the rpm of the take-up roller’s drive motor is proportionally decreased. This is to
compensate for the increase in linear velocity at the roller’s outer edge where the
sheet is coming onto the roll.
326 Chapter 10 Process Control Systems

We have discussed only five automatic control systems of various types in


this one area of paper processing. There are obviously many more that can be
found operating elsewhere in other parts of the manufacturing system. Some
control systems are so simple in their operation that during the design phase, the
designer probably didn’t take the time to analyze mathematically the system’s
behavior before having it built. In many cases, the system’s behavior isn’t all that
critical, or it can be easily adjusted through trial-and-error after being built.

10-6 LIQUID FLOWRATE SUPERVISION SYSTEM

This particular topic covers a lot of territory in controls design. Flowrate supervi¬
sion is used in controlling liquid flowrates or amounts in pipelines for chemical
supplying or foods operation, or perhaps for controlling hydraulic fluids in a
machine operation. Each installation of the automatic control system requires a
unique control design.
To begin with, let’s look at a typical flowrate control system shown in Figure
10-16. There are two important pieces of hardware in this system. One is the
servo-controlled valve used to vary the liquid’s flowrate in the supply pipe, and
the other is the flowmeter used to monitor the liquid’s flowrate. First, let’s look at
the servo-controlled valve. Figure 10-17 shows a diagram of a simple valve con¬
trolled by a servomotor. The valve’s seat is designed so that for a given amount of
rotation of the valve’s stem produced by the servomotor and its associated gear
transmission, a proportional flow of liquid is allowed to pass through the valve. In
some valve designs, provision has been made for a clutch to allow the motor to

Servo-controlled
Valve

Figure 10-16 A typical flowrate controller system.


Section 10-6 Liquid Flowrate Supervision System 327

Figure 10-17 A servo-controlled valve.

continue spinning if the valve seats completely or if the valve opens completely.
The clutch mechanism protects the motor from becoming overheated should ei¬
ther travel extreme be reached. In the case of very small valve designs, the clutch
is omitted and the motor is allowed to stall, having been designed to withstand the
increase in operating temperature resulting from the stalled condition. In either
case, there must also be a mechanical provision for accommodating the upward
and downward movement of the valve stem as the stem rotates. Some valve
designs use a linear motor rather than a rotational motor to move the valve stem.
Each valve design has its own unique advantages and disadvantages. For the
purpose of simplifying our discussion here, we use the valve design in Figure
10-17 to explain flow control. It’s probably the easiest design to follow and to
understand.
Our second important piece of hardware in the flow control system is the
flowmeter transducer. There are many designs to choose from for this applica¬
tion, one of which is shown in Figure 10-18. This particular model contains a
rotating turbine whose rotational velocity varies with the flowrate of the liquid
inside the pipe. A variable reluctance transducer picks up the passage of the
turbine’s blades and transmits a variable frequency (FM) AC wave whose fre¬
quency varies directly with the turbine’s speed. The FM signal is then demodu¬
lated to produce a variable DC voltage that is fed into a comparator for system
controlling.
328 Chapter 10 Process Control Systems

Output Signal

Figure 10-18(a) Schematic of turbine


flowmeter with installed sensor. (Dale
R. Patrick!Stephen W. Pardo, INDUS¬
TRIAL PROCESS CONTROL SYS¬
TEMS, copyright 1979, p. 122. Re¬
printed by permission of Prentice Hall,
Inc., Englewood Cliffs, N.J.)
(a)

(b)
Figure 10-18(b) Commercial turbine-type flowmeters. (Photograph courtesy of
The Foxboro Company, Foxboro, Mass.)

Another frequently used flowmeter design is shown in Figure 10-19. This


particular design utilizes a target that is suspended in the flow of the liquid or gas
in the pipe. The force of the fluid on the target causes it to deflect backward. The
amount of deflection is registered by a strain gage or other displacement-type
Section 10-6 Liquid Flowrate Supervision System 329

Figure 10- 19(a) Schematic of a targe:


flowmeter, iDale R. Patrick Stephen
U. Fordo. INDUSTRIAL PROCESS
CONTROL SYSTEMS, copyright 1979,
p. 77-. Reprinted by permission of
Prentice Hail. Inc.. Englewood Cliffs.
NJ.)

Figure 10-19(b) Commercial target-type flowmeter. Photograph courtesy of


Tee Eoxboro Company. Foxbow. Masse
330 Chapter 10 Process Control Systems

sensor which, in turn, produces either an analog AC or DC voltage output, or a


digital signal output if desired. In the case of our flow system in Figure 10-16, we
would want an analog DC voltage output so that it would be compatible with the
set point voltage at the comparator.
At this point, you may have recalled that we discussed a similar automatic
control system back in Chapter 9. In Section 9-4 we discussed a remote valve
positioner system that allowed a user to select a valve opening position that
controlled a particular flowrate. The position had to be selected by some prede¬
termined flowrate measurement. Once the valve opening position was selected, if
that flowrate ever changed, the selected valve opening would still remain con¬
stant. In other words, the system was not designed to compensate for variable
liquid flowrates inside the pipe. It was designed instead to position a valve stem
and to readjust itself to hold that position if some outside force caused that
position to change. Now, let’s compare this system to the one shown in Figure
10-16. The system in Figure 10-16 does not have the capability to set a particular
valve opening setting. Instead, for the system’s set point, the user inputs a de¬
sired liquid flowrate. The control system then opens or closes the valve; that is, it
self-adjusts until the desired flowrate is obtained. Notice that if the flowrate
should change in the pipe due to a change in liquid supply conditions upstream in
the pipe, the system can readjust or adapt its valve setting. The flowrate trans¬
ducer would sense this change and command the valve to readjust itself in order to
maintain the desired flowrate amount. This type of system is an adaptive-type
automatic control system. We talk more about this type of system later in this
chapter.
Now that we have become familiar with the hardware in our flow control
system, let’s go through its operation to make sure we understand how it oper¬
ates. Also, we can then better appreciate the basic difference between this adap¬
tive type of control system and the nonadaptive system in Chapter 9. Looking
again at Figure 10-16, the system’s operator sets the set point controller at the
liquid flowrate level he or she wishes to maintain in the piping system. The
controller’s dial face on the control panel would most likely be calibrated in
something like gallons per minute or cubic feet per second. However, in reality,
the controller would be outputting a DC voltage which would be proportional to
the dialed-in flowrate setting on the control panel. Let’s assume that our system is
just starting up with no flow in the piping system. With the set point data now
dialed in, the output voltage goes to the amplifier where it is amplified and pro¬
duces a proportional output power signal to the servomotor attached to the ser¬
vo’s valve. This amplified voltage causes the valve to open, allowing liquid to
flow through. The flowrate transducer then senses the flow and generates a pro¬
portional output DC voltage, which is sent back to the control system’s compara¬
tor. Depending on the amount of flow passing through the transducer, the magni¬
tude of its output signal is determined. This, in turn, determines the magnitude of
the error signal coming out of the comparator. The closer in magnitude the two
voltage values are coming from the set point controller and the flow transducer,
Section 10-7 Bottle Filling Process 331

the smaller the error voltage supplied to the servovalve. The servovalve will
continue to rotate until the error signal is zero; that is, until the transducer’s
output matches exactly the output of the set point controller. Notice what hap¬
pens if, because of the servovalve’s inertia, the valve opens too far beyond the
time when the error signal becomes zero (as could actually happen in some system
designs) and the error signal reverses polarity, causing the servomotor to reverse
its rotation. In other words, the servomotor will always try to correct itself until
eventually it settles at the proper flowrate setting. This is the same output charac¬
teristic we experienced in many of our other control systems, and so this system is
fairly typical. Hunting should be kept to within the typical amounts associated
with damping factors that are in the 0.4 to 0.7 range.

10-7 BOTTLE FILLING PROCESS

The food industry certainly has generated some very interesting challenges to the
automatic controls engineer and technician. Automating mixes and batches of
food ingredients can be extremely difficult where consistency is of utmost impor¬
tance. Figure 10-20 is a schematic of a bottling operation for soft drinks. In this

Figure 10-20 An automated bottle-filling operation.


332 Chapter 10 Process Control Systems

schematic, we see bottles arranged in a straight line on a conveyor system. In


reality, the bottles are usually on a circular rotating conveyor with an entrance
provision for inserting empty bottles and an exit for filled bottles. The bottles can
be filled by liquid level, by weight, or by a metered flow amount being supplied to
the bottles. The system shown in Figure 10-20 uses a liquid level detector sys¬
tem. In this system, a light source projects a beam of light through the bottle being
filled. This beam falls onto a photocell that produces a signal, which is sent to a
valve controller. The valve controller responds to this signal by opening the
supply valve over the bottle to be filled. When the filling liquid reaches a predeter¬
mined level inside the bottle, the light beam is cut off. The photocell, as a result,
turns off, causing the valve controller to turn off, which, in turn, tells the supply
valve to close. When the valve does close, the conveyor then automatically be¬
gins indexing. The indexing light source is then energized and the next bottle
moves into position under the filler tube. As soon as the indexing light beam is cut
by the next bottle, the conveyor is stopped. The filling light source is energized,
causing the next bottle to be filled and thus repeating the operation.
The important design problem to overcome in this particular operation is the
logic needed to tie together the two control operations so that they operate at the
proper times. You certainly don’t want the filling operation to be taking place
while the conveyer is moving, or vice versa. Also, provision must be made for the
likelihood that a bottle may be missing from the conveyor line. You will want to
be able to insert a control circuit that detects that possibility. You may also want
to be able to tell if the correct bottle is being filled. If you are running a 12-ounce
bottle operation and a 16-ounce bottle inadvertently got placed in the bottle line on
the conveyor, you will want to be able to recognize and cope with that situation.
This is where a vision system may be used, although simpler systems could be
devised to do the same bottle inspection. This depends on how radically different
the various bottle types may be that may appear on the conveyor lines. The
system shown in Figure 10-20 uses a Product Code Recognition or PCR laser
system. This system is virtually identical to the system often seen at the checkout
lanes of supermarkets for ringing up grocery prices on the cash register. The bar
code printed on each bottle is scanned by this system, and if the improper bottle
size data is detected, a pneumatic cylinder or ram ejects the bottle from the
conveyor line.
In place of the level detection system used for the bottle filling operation, a
weighing system may be used. As each empty bottle is positioned beneath the
filling tube, the bottle is made to be positioned on a small platform containing
either a strain gage or some other deflection-type transducer. This can be a linear
differential amplifier or an optically encoded displacement transducer. When the
proper signal voltage is generated by this device representing the proper filling
weight, a voltage comparator circuit then closes the filling valve and indexes the
conveyor to the next bottle.
Using a metered volume filling operation would be very similar to the system
described in Section 10-6. In this system, a flow transducer would be used to
Section 10-8 Telescope Tracking System 333

measure out a specified volume of liquid. Preferably, this transducer would be a


positive displacement type. That is, the transducer is designed to measure a
specific volume of fluid before discharging the fluid from its outlet. An example is
the rotating-vane meter (Figure 10-21). This is the type of meter used in gasoline

Figure 10-21 A positive displacement metering device used for filling opera¬
tions. (Dale R. Patrick!Stephen W. Fardo, INDUSTRIAL PROCESS CON¬
TROL SYSTEMS, copyright 1979, p. 127. Reprinted by permission of Prentice
Hall, Inc., Englewood Cliffs, N.J.)

pumps for metering dispensed gasoline. These transducers are extremely accu¬
rate. Since the exact volume of liquid that can be trapped between the rotating
vanes of this device is known ahead of time, it’s merely a matter of counting the
vane’s number of revolutions to determine the exact volume of metered liquid.
Accuracies greater than 0.1% are not uncommon for this device. Knowing the
number of revolutions ahead of time needed to produce a certain volume, the total
revolution count can be converted to a DC voltage and sent to a comparator
circuit for controlling the filling valve. This voltage conversion may be easily
handled by a digital-to-analog converter.
The filling operation, the conveyor indexing system, and the bad bottle
ejection system may all be thought of as bang-bang servo systems, since they
behave much like the one we analyzed in Section 9-8. That is, these control
systems are either all the way on or all the way off. This is a fairly typical
characteristic of the many control systems used in process control applications.
They are generally cheaper than the more complex linear servo systems and are
easier to maintain.

10-8 TELESCOPE TRACKING SYSTEM

This next example of a process control automatic control system is a fairly recent
innovation in astronomical research. It is a unique control system and has saved
labor and increased the accuracy of telescopic tracking.
334 Chapter 10 Process Control Systems

We are all aware that the earth rotates daily on its axis creating the illusion
that the sun, the stars, and the planets rotate around the earth. Whenever the
astronomer wishes to photograph a portion of the sky using an optical telescope,
lengthy time exposures are required. These time exposures may run several
hours, and in some instances, even days. The reason for these long exposures is
to gather enough light in order to record an image. Many of the objects that are
photographed are hundreds of thousands or millions of light years away, and by
the time that light reaches earth, it is understandably quite weak. (One light year
is approximately 5.87 trillion miles.) What we may not be aware of, however, are
the intricate mechanisms needed to keep the telescope with its installed camera
film pointed at the object or objects being photographed. These mechanisms es¬
sentially allow the earth to rotate beneath the telescope while the telescope re¬
mains fixed, continuously pointing at the desired object. While this description is
somewhat flawed since the earth really doesn’t rotate beneath the telescope,
because the scope is rigidly attached to it, it is partially true. The way most
telescopes are presently mounted on their earth-fastened mounting bases allows
this to happen. These mounts are called equatorial mountings. In order to appre¬
ciate the control system we study next, it’s necessary to understand these mount¬
ing systems.
Figure 10-22 shows a telescope mounted on its equatorial mount. Because
all objects in the sky appear to rotate around a point located close to the north star
(since this is the point in the sky where the earth’s rotational axis is pointed), it
becomes a matter of designing a two-axes mount, one of which points at the north

I Polaris
—y (the North
I Star)

Observing
Area

Figure 10-22 The motions of an equatorial-mounted telescope.


Section 10-8 Telescope Tracking System 335

star. This axis is called the scope’s polar axis. A second axis, called the declina¬
tion axis, allows the scope to swing away from the equatorial axis and point to any
other spot in the sky. In order to track an object automatically at any of these
points, it becomes a matter of attaching a drive motor to the polar axis to allow the
scope to slowly rotate about this axis. Of course, the motor’s speed must be
precisely synchronized with the rising and setting time of the objects in order to
keep up with their movement. This motor is called the scope’s clock-drive. While
this type of system works acceptably well, it is still an open-loop control system
that is vulnerable to outside influences—influences such as accidental bumping of
the telescope frame by operating personnel, variable refraction of the object’s
light reaching earth due to sudden variations in the earth’s atmospheric density,
and rapid motion of the object in the sky due to its being near the earth or because
of very high orbital velocities. All of these variations can be compensated for
using present control designs, but not without requiring complex electromechan¬
ical system designs and requiring constant, almost minute-by-minute, supervi¬
sion. Typically, a technician is required to make small minor touch-up correc¬
tions to the clock-drive using what is called a slew motion pad. He or she uses a
smaller guide scope, similar to a target scope of a rifle, to follow a bright star, or
guide star, which is usually located near the photographed object. This guide
scope is attached to and is parallel with the main frame of the larger scope. In
other words, the scope is slewed in whatever direction is necessary to bring it
back on target should either it or the object wander off.
The following system is a simplified version of a closed-loop automatic
tracking system now being used in some observatories around the world (Figure
10-23). It is comprised of two sets of photocells mounted in the guide scope.
Each pair of photocells has its outputs sent to a differential amplifier, power
amplifier, and motor. A second closed-loop system, which is a positional system,
is also shown which is used for initially pointing the telescope at the correct
object. The entire system works like this: The object to be photographed is often
invisible to the scope’s operator. However, its coordinates are usually precisely
known. (These coordinates are like the coordinates of an object located here on
the earth’s surface where two x-y coordinates are needed to form an intersection
denoting where the object is located.) A guide star is then selected that’s known to
be near the object of interest. That object’s positional coordinates are then en¬
tered into the computer’s control system. The computer corrects the coordinates
for the time of day (converting the right-ascension coordinate to what is referred
to as the object’s hour angle) and sends that information to the scope’s two setting
circles. This information is compared to the existing setting circle values and, if
an error voltage is present due to different setting circle values, the scope is
, automatically moved by means of that error voltage until the error voltage drops
to zero. This is done on both axes. Because of the massiveness of the scope, little
if any hunting has been designed into the system. Damping factors are usually
fairly large for this size of system. Response time is also quite slow. Once the
scope is positioned on its target, a guide star is found and centered on the guide
336 Chapter 10 Process Control Systems

scope’s cross-hairs. At this point, since the opposite photocells are receiving
equal light from the guide star, the comparators for each photocell pair will be
generating no error signals. However, should the guide star be off center from the
cross-hairs only slightly, one of the photocells will receive more light than the
other. Consequently, an error signal is generated in the otherwise balanced com¬
parator, and the appropriate drive motor is energized to bring the guide star back
on center or until the error once again drops back to zero voltage. Most likely,
during an automatic correction both comparators will be affected.
It is interesting to note that some telescope drive manufacturers have done
away completely with the polar axis. Instead, they have mounted the scope on
what is called an altazimuth mount. This type of mount is very simple to con¬
struct and is similar to the mount used on surveying equipment. Here, a computer
continuously drives both what are now the altitude axis and the azimuth axis of
the scope in an effort to keep the scope on its target. This is a somewhat revolu¬
tionary design and has worked out very successfully. It eliminates the difficult
task of constructing a polar axis that is exactly parallel to the earth’s rotational
axis. Automatic slewing works very well with this system and is almost manda¬
tory, since the polar axis is no longer used.

10-9 DIE CASTING PROCESS CONTROL SYSTEM

Die casting is a very old and interesting process. While the process has been
around for a long time, it has only been since the 1960s that the die casting
industry has developed an automated process for producing castings. And it was
Section 10-9 Die Casting Process Control System 331

as recent as the early 1980s that it was proven that molten aluminum, a material
often used in die casting, required a certain insertion velocity into the die in order
to produce a good die-cast part. Let’s look at a typical die casting machine and
see how it is presently being automated.
Figure 10-24 shows a typical die casting machine and all its associated
parts. What we are particularly interested in is referred to as the machine’s shot
end. This is the end where the molten metal is ladled in by a robot and is then
“shot” or rammed into the die. The robot ladles a measured amount of molten
aluminum into the pour-hole (not shown in Figure 10-24) of the machine’s cold
chamber. This is similar to the loading of a cartridge into the breach of a cannon,
except in place of a solid cartridge material being inserted, we’re inserting a liquid
material. After the ladling, a hydraulic ram immediately shoots the metal through
a hole in the die casting machine’s front plate and into the die cavity, lo¬
cated inside the two die-halves situated between the front plate and the ma¬
chine’s traveling plate. The traveling plate is clamped against the die halves and
front plate by means of a mechanical toggle system that can generate tre¬
mendous clamping forces anywhere in the range of 200 to as much as 3000 tons,
depending on the size of the machine. The toggle system, during this clamping
procedure, pushes against the machine’s back plate. The back plate in turn
stretches the four tie bars (only two are seen in Figure 10-24) that are fixed to the
front plate. This stretching causes the two die-halves to be further “squeezed”
together between the traveling plate and the front plate while the “shot” is
being made. After a short curing or freezing time, the toggles are released
and the traveling plate moved back, causing the die-halves to part. A second
robot then reaches in between the two die-halves and extracts the newly made
part.
Now that we have discussed the overall action of the die casting process,
let’s look at some of the automation techniques being used. First, let’s go back to
the machine’s shot cylinder. Earlier, we suggested that the molten aluminum
being shot into the die must be delivered to the die at a particular velocity by the
shot cylinder, or ram, in order to make a good part. This velocity is called the
critical velocity. To obtain and hold this critical velocity, we have to employ a
method that enables us to monitor and to adjust automatically the shot cylinder’s
ramming velocity to maintain this velocity. To monitor the velocity, an optical
encoder can be attached to the shot cylinder’s piston rod to measure its forward
speed. Since the optical encoder generates digital information in the form
of pulses, this can be converted to a DC voltage and compared to the desired
dialed-in velocity, which is the desired critical velocity. If an error voltage is cre¬
ated at the comparator, this voltage can readjust a hydraulic proportional valve
to either decrease or increase the shot cylinder pressure for correcting the
next shot. Analog-type transducers can also be used in place of the digital
optical encoders; however, using digital signals wherever possible within the
system will make the control system’s signals as immune to electrical noise as
possible. Unfortunately, electrical noise is extremely prevalent around many
338
Figure 10-24 Die casting machine. (Wayne AlofslJames R. Cars tens, MECHANICAL
MAINTENANCE AND EVALUATION OF DIE CASTING MACHINES, copyright 1987.
Reprinted by permission of the Society of Die Casting Engineers, River Grove, III.)
Section 10-9 Die Casting Process Control System 339

die casting shops because of electrical welding and heavy reliance on electro¬
mechanical relays.
Now, let’s look at the so-called closing end of the die casting machine. This
is the opposite end of the machine where the rear end of the toggle system is
anchored to the machine’s back plate. In order to monitor the amount of force
being generated within the tie bar system and to insure that all four tie bars are
evenly loaded, we need to install transducers that can sense deflection in the tie
bars. The logical choice would be to use strain gages; however, it is difficult to
maintain calibration with strain gages in an industrial atmosphere, not to mention
their temperature susceptibility (of which there is plenty in a die casting shop).
The linear variable differential amplifier or the optical encoder would be good
choices instead. The optical encoder produces a digital signal and would have
that particular advantage over a linear variable differential amplifier system. Both
systems are shown in Figures 10-25 and 10-26, respectively. With these systems,
a particular locking tonnage can be dialed into a comparator and the machine then
locked up on a die. The stress in each of the four tie bars can then be compared to
one-fourth of the inputted desired tonnage. The outputs of the four tie bar indica¬
tors can then be compared to the dialed-in data and the machine adjusted accord¬
ingly. This adjustment is performed by either tightening or loosening the appro¬
priate tie bars to obtain the desired balance loads. This tightening or loosening

Figure 10-25 Using the LVDT as a stress indicator. (Wayne AlofslJames R.


Cars tens, MECHANICAL MAINTENANCE AND EVALUATION OF DIE
CASTING MACHINES, copyright 1987. Reprinted by permission of the Society
of Die Casting Engineers, River Grove, III.)
340 Chapter 10 Process Control Systems

Light Source

Figure 10-26 Using the optical encoder as a stress indicator. (Wayne Alofs/
James R. Carstens, MECHANICAL MAINTENANCE AND EVALUATION OF
DIE CASTING MACHINES, copyright 1987. Reprinted by permission of the
Society of Die Casting Engineers, River Grove, III.)

process can be fully automatic and can be done by energizing a gear drive system
that screws the tie bars into or out of threaded mounts. One or all the tie bars can
be moved, depending on which tie bar requires this adjustment.
Much development is needed yet in automating die casting operations. This
is a relatively new field and is of particular importance to the automotive industry,
since they depend heavily on the die casting process for the manufacturing of their
components.

10-10 A PROPORTIONAL PNEUMATIC CONTROLLER

Proportional pneumatic controllers are used to create proportional pneumatic


signals to actuators requiring intermediate positioning capabilities. An example
would be the controlling of a damper setting on an air conditioning system, where
the damper would have to report to an infinite number of intermediate positions
for mixing air in the duct work depending on the heating or cooling demands of the
system (Figure 10-27). The damper’s setting would be controlled by an attached
pneumatic cylinder whose plunger extension would be controlled by a propor¬
tional pneumatic controller.
Proportional pneumatic controllers have, to a great extent, been replaced
with electronic proportional controllers. However, there are numerous places in
industry where there is still demand for them. Pneumatic controllers are used in
Section 10-10 A Proportional Pneumatic Controller 341

Figure 10-27 A proportional damper


mixing system in an air conditioning
system.

volatile atmospheres where operating electrical devices would be too dangerous.


Also, the cost of one of these controllers is considerably less than a comparable
electronic controller.
Figure 10-28 is a schematic of a typical proportional pneumatic controller
being used in conjunction with a pneumatic temperature sensor. The system
works like this: A temperature sensor is installed in a room having ductwork that
supplies both heated and chilled air to that room. The amount of heating or
cooling air is determined by a damper blade that is actuated by a pneumatic
cylinder. If the blade is caused to rotate counterclockwise, more cooling is al¬
lowed into the supply ducts. If the blade rotates clockwise, more heating is sup¬
plied to the supply ducts. The temperature sensor is a bimetallic temperature
element attached to an air flow control valve. If the temperature happens to be
rising, the bimetallic strip expands, therefore restricting the airflow valve and
allowing less air to flow into the bellows. A falling temperature causes the bime¬
tallic strip to contract, therefore opening the air valve and allowing more air to
flow into the bellows. The desired temperature, or set point controller, is a spring-
loaded screw adjuster. The system is calibrated so that when the bellows is ex¬
actly offset by the spring-loaded screw adjustment, that particular temperature
exists in the supply duct. Let’s assume now that the temperature rises above the
set point. This will cause less air to enter the bellows, causing the blocking lever
to move left in the diagram and blocking more of the air-bleed port. This, in turn,
causes more air to be bypassed into line A. At the same time, though, more air is
also bypassed into line B and into the fluidic amplifier. The amplifier’s diaphragm
expands, pushing the actuator control valve to a more closed position. The actua¬
tor cylinder causes the piston to move downward in the diagram which, in turn,
causes the damper to rotate counterclockwise, thereby increasing the cooling.
Let’s assume now that the temperature sensor senses a falling temperature
in the supply duct. The falling temperature causes the bimetallic strip to contract,
342 Chapter 10 Process Control Systems

therefore opening the air supply valve and causing the bellows at the blocking
lever to expand. This causes the blocking lever to swing right in the diagram.
This creates a greater bleed-port bypass. As a result, less air is delivered to the
fluidic amplifier through line B and less air also to be delivered by line A to the
actuator control valve. The fluidic amplifier’s diaphragm now relaxes, therefore
opening the actuator control valve further. More air will now flow to the actuator,
causing the piston to move upward in the diagram and moving the actuator clock¬
wise. This action increases the heating supply to the supply duct.
This type of closed-loop temperature control is very effective. It's inexpen¬
sive and quite rugged. Its only major disadvantage is having to have a continuous
air supply along with extensive piping. Periodic maintenance is needed to assure
a leak-free system.
Section 10-11 A Remote Control Antenna Rotator 343

10-11 A REMOTE CONTROL ANTENNA ROTATOR

The final process control system that we discuss here is a remote control antenna
rotator system. This is a system used to enable an operator to position a commun¬
ications antenna from a remote location so that the antenna can be directed at a
distant receiving antenna. With the closed-loop servo system, the antenna will
self-correct itself if gusts of wind or any other disturbances should happen to
attempt to move it off course. Figure 10-29 shows a diagram of this system. The
system uses a CX-CT 400 Hz synchromechanism for the input command signal
and checking the output position of the antenna. This is a very precise system and
has many applications in other fields of control. It is quite similar to the remote
valve positioner example discussed in Section 9-4. The reason for the 400 Hz
CX-CT system is only because this system was used frequently in aircraft con¬
trol. The 400 Hz frequency is used extensively aboard aircraft and by the military
in general.
Taking a look at the system, we see that a DC motor is used for turning the
antenna. The reason for this is that DC motors are capable of much higher torque

Amplifier

Figure 10-29 An antenna rotational controller.


344 Chapter 10 Process Control Systems

outputs as compared to AC induction motors. The disadvantage in using this type


of motor, however, is that some means must be used to convert AC to DC to make
the motor control compatible with the AC CX-CT system. This is the reason for
the demodulator. The output of the CT is demodulated so that its DC equivalent
voltage can be eventually compared with the DC output of the tachometer me¬
chanically coupled to the antenna’s drive motor.
This is how the system works: It must be assumed that the system has been
initially calibrated. That is, whatever bearing heading is dialed into the CX back
at the control station, the antenna system located at its remote site is guaranteed
to be pointing at that same bearing heading. In other words, there is zero error
signal being generated within the control system. Initial calibration would proba¬
bly involve a rather simple procedure similar to placing the CX at 0° heading and
then adjusting antenna so that it points due north, 0°. To analyze the system’s
operation, let’s assume that the operator adjusts the CX for a compass bearing of
30° east of north. The antenna, however, is pointing at, say, 150° from a previ¬
ously commanded heading. (All compass headings in this example are measured
clockwise from north.) Because of the difference in headings of the CX and CT
units, an error signal is generated at the CT in the form of a certain amplitude AC
waveform that is either in phase or out of phase with the supply voltage AC
waveform. Remember that AC control systems always use the supply voltage
waveform for referencing phase relationships. In this case, the system is designed
so that out-of-phase signals at the CT produce a counterclockwise rotation of the
antenna. The amplitude of the error signal causes the drive motor to rotate at a
proportionally related speed. The higher the amplitude, the faster the rotation. In
this example, an out-of-phase signal will be produced that will cause the servosys-
tem to rotate the antenna counterclockwise. There will be a relatively large ampli¬
tude error signal that will be produced due to the differences between the set point

Antenna

Input
(Rad.)

Figure 10-30 Block diagram of antenna system shown in Figure 10-29.


Review Questions 345

signal and the original resting point of the antenna (150° southeast). To decrease
the likelihood of over-shoot and to dampen an otherwise lively system because of
the components being used, a tachometer is used to convert velocity into posi¬
tion. This system’s block diagram is shown in Figure 10-30.

SUMMARY

The systems presented in this chapter are but a very few of the many workable
systems that are being used today. The ones described here have been simplified
so that their basic operating concepts can be better seen. Despite these simplifica¬
tions, if you understand what has been presented in this chapter, you will have
little difficulty understanding the more detailed systems encountered in industry.
If there are concepts that you are still struggling with in understanding automatic
control systems, reread the earlier chapters. Chapter 10 represents a parting from
the theoretical and an entry into the practical real world systems. Unless the
theory is adequately “nailed down” at this point, you may have difficulty continu¬
ing with the practical information from this point onward.

REVIEW QUESTIONS

10-1. What advantage does an analog automatic control system have over a
digital automatic control system?
10-2. What two advantages does a digital automatic control system have over an
analog automatic control system?
10-3. Why was it necessary to develop the gray code to replace the binary code
in certain encoding situations?
10-4. What are the three major types of power sources used in robotics today?
List advantages and disadvantages of each.
10-5. Explain what is meant by the term bit map as applied to optical control
systems.
10-6. Explain the function of a digital comparator circuit. What are its similari¬
ties and differences when compared to an analog voltage comparator or
differential amplifier?
10-7. Explain what is meant by the term adaptive control in a process control
system.
10-8. Explain the purpose of a tachometer as used in an automatic control
system.
10-9. What would be the advantage of using an optical encoder for the measuring
of very small increments of displacement versus using a linear variable
displacement transformer for the same application?
10-10. Why is adaptive control so important in process control systems? In other
words, why would you want to use it in a process control design?
346 Chapter 10 Process Control Systems

REFERENCES

Hunter, Ronald P., Automated Process Control Systems, Englewood Cliffs,


N.J.: Prentice-Hall, Inc. 1987.
Johnson, Curtis D., Process Control Instrumentation Technology, New York,
N.Y.: John Wiley & Sons, 1977.
Patrick, Dale R., Industrial Process Control Systems, Englewood Cliffs, N.J.:
Prentice-Hall, Inc., 1985.
Potvin, Jean, Applied Process Control Instrumentation, Reston, Va.: Reston
Publishing Company, Inc., 1985.
11-1 THE PURPOSE OF SYSTEM COMMUNICATIONS

Up to this point, we have more or less taken for granted the fact that control
signals, by some means or another, traveled trouble-free from one portion of a
control circuit to another. We gave little thought to the means or methods used in
transmitting these signals between circuits. We assumed that they arrived at their
destination intact and unaltered. In this chapter, we study the various methods
used in conditioning the control signals for travel and why this preparation is
done. There are definite advantages to using certain means of preparing signals
for transmission and for using certain decoding methods on the data that are
imbedded in these transmissions.
Some of the discussion here involve a light dosage of radio theory and
computer data transmission. To make the discussion a little easier to digest, it has
been broken down into three major categories:

1. analog data transmission,


2. digital data transmission, and
3. carrier data transmission.

Each category is discussed, and advantages and disadvantages are presented in


each. In reality, carrier data transmission techniques may be considered an ana¬
log form of communication. As a matter of fact, as we soon find out, carrier
transmissions are used primarily to transmit analog-type data. However, in an
effort to make the discussions a little easier to understand and to be more concise,

347
348 Chapter 11 System Communications

the carrier data transmission discussion was separated from the analog data trans¬
mission material.
Perhaps you are asking yourself, why not just send the output of one circuit
directly into the input of the next? Why bother to transform the output into some
form of data signal before transmitting it to the next input? Unfortunately, things
aren’t quite that simple in hooking circuits together. A problem that could de¬
velop with the direct hook-up of circuits is this: The transmission of original
variable voltages or currents is fine for small distances measured in, say, inches or
a few feet. But when you are talking about hundreds of feet or miles, there are
problems with outside electrical noise or interference, not to mention the prob¬
lems with DC or AC voltage losses due to resistance in the wires. In addition,
there are problems with lack of compatibility between circuit component inputs
and outputs. One may require an AC voltage while another may require a DC
current. There are also problems with impedance matching between these inputs
and outputs in order to obtain maximum power transfer between circuits. All of
these factors can really complicate an otherwise simple direct hookup. Therefore,
in this chapter we attempt to unravel some of the mystery and jargon used in
describing the various data communications systems used in automatic control
systems.

11-2 ANALOG DATA COMMUNICATIONS

In the beginning days of automatic control systems, analog data communications


was the only practical means of sending signals from one portion of the system to
another. However, as time passed and control systems became much larger and
more sophisticated, other means of data transmission had to be contrived. A good
example of this problem is the remote antenna rotator system described in Section
10-11. It’s conceivable that in this installation the antenna could be located sev¬
eral thousand feet from the control station as depicted in Figure 11-1. This is a
common situation where the separation is necessary to reduce the problem of
radio frequency interference (RFI) in the control station, especially if the antenna
should be pointing directly at the station. In addition, to further reduce this prob¬
lem and to increase the range of the antenna’s radio transmissions, the antenna
may be mounted on top of a tower hundreds of feet high. This adds to the path
length of the controlling signals and to the control cable itself. Using an analog
control system for this application would be somewhat difficult. However, we
discuss a possible solution in a later section.
Analog systems (Figure 11-2, for example) are comprised mainly of those
system components whose outputs are a variable amplitude voltage or current.
The amplitude of this current or voltage is a linear representation of that compo¬
nent’s input. Many components, such as transducers, synchromechanisms, am¬
plifiers, signal conditioning circuits, etc. are designed so that their inputs and
outputs are relatively compatible; that is, the output of one can be fed to the input
Section 11-3 Carrier Modulated Control Systems 349

I I
i-1
Buried Control Cable

Figure 11-1 Typical communications antenna installation.

of the other with only minor circuit modifications. Many process control devices
are designed to put out a current in the range of 4 to 20 mA. In other words, many
circuits are designed so that 4 mA represents the lowest expected output signal,
while 20 mA represents the highest expected output. Some form of circuit scaling
is needed to insure the meaning of this range. Obviously, this very low current

Figure 11-2 A process control system using a 4-20 mA analog circuit automatic control
components.
350 Chapter 11 System Communications

range cannot be expected to drive very many output devices by itself. It would
first be necessary to boost the power by using a power amplifier.
Figure 11-2 is an example of an analog-type automatic control system. The
humidity sensing element used to detect the proper amount of moisture content in
a packaged cookie product has a 0-1 mA output. This output signal is sent to a
signal conditioner where it is amplified and the sensing signal “scrubbed” of any
extraneous electrical noise that might create false data. This signal, which is now
in the 4-20 mA range as a result of amplification, is sent to a power amplifier
where the current signals are converted to an analog drive voltage to drive the
cookie’s conveyor motor. Because of the relatively short control lines needed in
this system, the analog signals may be transported by ordinary wire or cable,
although shielded cable is highly recommended. (See also the discussion in Sec¬
tion 12-7.1.)

11-3 CARRIER MODULATED CONTROL SYSTEMS

Carrier modulation (sometimes referred to as CW, meaning continuous wave) is


seldom used for transmitting data in control circuits. It was used extensively
many years ago, but not for this purpose. It was used instead for the transmission
of Morse code for communications purposes. The transmitter’s carrier was inter¬
rupted to form combinations of dashes and dots, each group forming an alphanu¬
meric character or punctuation. It is still used in limited applications for the
transmitting of data strictly for informational purposes, such as in emergency
communications, but not for circuit control. There are applications, though,
where the systematic turning on and off of a transmitted carrier is used in the
controlling of bang-bang servo systems, but these cases are somewhat limited. An
example would be the controlling of a garage door opener system. A hand-held
transmitter is turned on momentarily to energize a receiver at the garage door.
The receiver, when receiving this transmitted signal, turns on a motor which lifts
the door. The door stops when limit switches installed in the door's track path
detect the door’s maximum travel. Another short burst of transmitted carrier
from the transmitter then reverses the procedure and closes it.

11-4 AMPLITUDE MODULATED CONTROL SYSTEMS

Amplitude modulation data communications refers to the transmitting and receiv¬


ing of radio frequency signals that contain data represented by the varying ampli¬
tude of the carrier signal. The transmission and reception of AM radio and televi¬
sion signals are two good examples. In the case of television, the data being
transmitted by amplitude modulation is the color information for the TV picture.
The changing amplitude of the carrier represents the varying intensities of each
frame or picture being transmitted. In the case of AM radio, the sound is trans-
Section 11-4 Amplitude Modulated Control Systems 351

mitted by the same means. The sound’s amplitude and pitch is represented by the
amplitude and frequency of the modulation “envelope” surrounding the carrier.
This is demonstrated in Figure 11-3.

Carrier Frequency

Figure 11-3 An amplitude-modulated (AM) carrier wave.

To understand how an amplitude modulation data system works, let’s refer


to the block diagram for the remote antenna controller described earlier in Figure
10-32. Let’s assume that we want to adapt this system to an AM control system
having the system measurements shown in Figure 11-1. The distance between the
CX and the remaining control system components, which are all located in a
control shed at the antenna tower site in this figure, is 2,000 + 250 feet. It should
become obvious that transmitting low-amplitude voltages in an analog system may
result in too-weak control signals by the time the transmitted signal voltages
appear at the tower’s base. Long lengths of transmission lines tend to be suscepti¬
ble to stray voltages from outside sources that are inductively and capacitively
coupled into the line. Using shielded cable reduces the problem but doesn’t elimi¬
nate it entirely. The result is gross distortion of the control signals because of
these stray induced voltages.
We use an amplitude modulated carrier system. This is a system borrowed
from radio frequency transmission technology and involves transmitting a radio
signal at a specific radio frequency. The transmitted signal is comprised of an rf
carrier that is amplitude modulated by the CX’s three output phase voltages. We
have each of the three control phase voltages vary the amplitude of three transmit¬
ted tones, called subcarriers, having three different frequencies of, say, 400 Hz,
730 Hz, and 960 Hz. We could have picked any three frequencies as long as they
are sufficiently separated from each other. These three values were picked,
though, because their harmonics and subharmonics will cause a minimum of
interference with each other. Each of these tones will then modulate the transmit¬
ted carrier frequency signal, as seen in Figure 11-4. In our system, the 400 Hz
control voltage line frequency (or whatever other power supply line frequency
happens to be used, such as 60 Hz) is removed through rectification. In other
words, the AC control signal is converted to a DC voltage whose varying ampli-
II

ll 05
> c
CJ
> c >•
o CD CD (D
c D
<D >
CJ *4— CD
D CD 4-1
cr L_ O 7s
CJ
a; LL
i_
>>
Ll- U o
CD
CD c CJ
4-»
c CD § Q
o D cj
H “O 0)
*—
o LL

Transmitting control signals using AM.


Figure 11-4
60 or 400 Hz

352
Section 11-5 Frequency Modulated Control Systems 353

tude coincides with the varying amplitude of the AC signal. The DC signal is then
allowed to modulate, that is, vary the amplitude of one of the audio tones. Then,
the audio tone with its varied amplitude is allowed to modulate an r.f. carrier. The
AM (amplitude modulated) carrier, as it is now called, is of very low power,
typically less than 500 milliwatts. Not much power is required to transmit a signal
down the 2,250-foot path of cable. Besides, this reduces any likelihood of inter¬
ference with other services in the area of the system. The frequency of this carrier
is also usually quite low compared to normal communications frequencies, typi¬
cally in the area of 10 to 500 KHz. The reason for this is to discourage radiation of
the signal through the air over long distances and to discourage interference with
the much higher communications frequencies normally being'used by the other
services. Since shielded cable is used for control signal transmission and because
the cable is buried in the ground for the most part, little radiation will be lost from
the cable.
Each of the three phase control signals from the CX are now converted to
audio tones and transmitted simultaneously down the control cable to the waiting
control system at the antenna tower base. We need a means of deciphering the
voltage and phase information being transported by the modulated carrier. There¬
fore, a radio receiver must be used to demodulate the carrier and “strip off” this
information. If this receiver had a speaker attached, you would hear a chorus of
three amplitude-varying tones, but they would still be of no use to the control
circuit at this point. Specially tuned filter circuits are needed to separate the tones
back into three separate discrete channels, just as they were back at the CX.
Then, following this filtering process, they must be mixed with the original control
voltage frequency (400 Hz in our case) to get them back into their original forms
for sending on to the CT. Figure 11-5 shows the complete remote control system
now modified for the automatic controlling of our antenna.

11-5 FREQUENCY MODULATED CONTROL SYSTEMS

Figure 11-6 shows an example of a frequency modulated control signal. Instead of


modulating the amplitude of a transmitted carrier as we did in Section 11-4, we
modulate the carrier’s frequency. In other words, again using the example of the
antenna rotator system, the only changes we make to our system’s design are to
change the type of modulator used on the transmitter and the type of receiver used
at the tower site. Both the transmitter and receiver must now be designed for
frequency modulation and demodulation, respectively.
The advantage of frequency modulation over amplitude modulation of con¬
trol signals is that FM is much more noise immune as compared to AM. Very
little man-made or nature-made electrical noise is FM. It’s for this very reason
that FM is quite popular for commercial broadcast applications.
There is some room for debate here as to whether or not a transmitted carrier
is actually needed to send and receive tone encoded information. Why not simply
354 Chapter 11 System Communications

Figure 11-5 Complete remote control antenna controller system.

transmit the tones themselves without an rf carrier, must like a transmitted tele¬
phone signal? It can be argued that audio tone frequencies are nothing more than
very low frequency radio waves having different propagation characteristics as
compared to the higher frequency carriers. Experimentation is still being con¬
ducted on which frequencies propagate better over lines having lengths similar to
the example we discussed here. One advantage to simply transmitting the tones is
that you won’t be troubled quite as much with the signal leaving the wire. The
wire has less of a tendency to act as an antenna as compared to the case of the rf
carrier transmission setup. Another advantage is that the circuitry is somewhat
less complicated. Figure 11-7 shows a schematic of a simple but effective means
of converting a variable amplitude AC voltage to a variable frequency tone. The

Figure 11-6 A frequency-modulated (FM) carrier wave.


D

+ 12 V

S-* ^
co
6C• PQ
tq
o .S U
o "O 0^
> o
(j
ca> O
o
s-
*-1 ^ £
C O ^
o
o s
<uc -S3
• >»»*0«
o S ?s
§■
x>
£
Vh
<u
-t~<
s-
a>>
c *8
o g
o
>, £ ^«
uc 'o
a> <D3
Cl
cr
<v
^ +3
h.
E
< o <3X
<u s:
3 o ?S
OX)

■i—>
• '*«*
>3 -S3
o> >3
+3<X
_a>
3 CX +3&a
-S3

•S
’C(3
_5< -S3
3
>
CX
"3!
'Ki

* *•S3v.»» =QX>
v-

=X
3—

ha
S
WD
E* G -S3

oa)
cc

355
356 Chapter 11 System Communications

circuit is comprised of four main subcircuits. The first circuit converts the incom¬
ing variable AC voltage signal from the CX into a variable amplitude DC signal.
This, in turn, is sent to an amplifier to increase its amplitude to a point where it
will drive a voltage-to-frequency oscillator (or VCO). This circuit converts the
DC signal into a variable frequency. From there, the signal is sent to an amplifier
whose job is to boost the line’s transmitting power to prepare the signal for
transmission down the control cable to the receiving site. Virtually any type of
low-power amplifier in the range of a watt or so would work for this application.
About the only thing that is somewhat critical about this circuit is the matching of
the control line’s impedance to that of the amplifier’s output. Even that is not a
difficult job, since what minor mismatch may occur at this point can be compen¬
sated for by merely increasing the gain of the amplifier somewhat.
The decoding of the received tone at the opposite end of the control line is
equally fairly simple. Figure 11-8 illustrates a schematic of such a system. A
Schmitt trigger is used to “strip-off” any line noise in the control cable. The 555
IC circuit does the actual variable frequency decoding, whereas the field effect
transistor (FET) produces a very stable and reliable output voltage that is inde¬
pendent of any fluctuations in the circuit’s supply voltages. The DC-to-AC in¬
verter converts the output variable DC voltage to an AC voltage whose frequency
is 400 HZ once again.

11-6 PHASE MODULATED CONTROL SYSTEMS

The discussion of phase modulation is presented here only to complete the discus¬
sion on modulation techniques. There is little advantage in using this form of
modulation over that of FM. One possible advantage is the somewhat better noise
immunity as compared to FM. The method of detecting a PM (phase modulated)
signal is very similar to that of FM methods. Essentially, phase modulation is a
method of varying the phase between the transmitted carrier and the modulating
data. The modulating data is in the form of a frequency-varying AC signal. The
phase variations are then detected by a receiver which, for all practical purposes,
can be an FM-type receiver.

11-7 DIGITAL DATA COMMUNICATION

Digital signals are in the form of pulses or square waves. They are either transmit¬
ted in groups to form coded alphanumeric data or they are transmitted in continu¬
ous but variable numbers allowing their frequency to be counted for the purpose
of deciphering the data (Figure 11-9).
With the advent of the microcomputer in the early 1970s, it became evident
that control circuits would also be affected. Circuits were redesigned so that
direct interfacing with the computer would become possible. This then would
N
X
o
o

osi -Cl

>5 “a
-c<<
• »»a
o
-c 3
• ■•**
£ s:
£C OJt

D
o
a>u -c>^
b.
CJ
l.
<DN £ ^
c O
JDCD o O
4-J
CO ^
<u F3
tq
£P U
B <q
o £
> O
>5
^ ^
S g ^
O' ^ -o
<u ^ .°0
£ O V.

CJO q §-
c cq Ci
'-a ^
o ^
oa> 6
cq U
•a tq -
a> ^ °3
c -5; 05
2 £
< <n
'n 02
•n
ac tq ^
H: -xs
a> K
Jfa c
_ 5 £
.gf> o £
fa ^ q
+5 V O

2 1
2 &

357
358 Chapter 11 System Communications

Variable Frequency Pulses

Equivalent FM
Pulse Modulation
Signal
■>- Time

Figure 11-9 Variable frequency pulse transmission.

allow the computer to “talk” directly with the circuit components with no inter¬
mediate data conversions necessary. But there were other reasons for going to
digital signal control. Digital signals were far more immune to electrical noises
than were some analog signals. By far the worst affected control signal was the
analog signal. As mentioned earlier, most man-made and naturally-made electri¬
cal noise is amplitude modulated. Consequently, the receiving circuits that were
designed to respond to this type of transmission could became easily confused
with this noise. Little such noise problems existed with digital-type circuits, how¬
ever. But it was soon discovered that because of increased complexities of cir¬
cuits, increased gain, and because much lower voltage values were being used for
output signals, the digital circuits become more susceptible to ground loops and
stray r.f.-type noise. (Ground loops are undesirable electrical currents flowing in
the ground and through wire commons or cable shielding, behaving like uncontrol¬
lable feedback signals, producing undesirable electrical circuit noise in systems.)
Solid state components, unfortunately, have a habit of acting as miniature radio
detectors in that they often and inadvertently pick up radio transmissions. But
despite this major flaw, there was an added bonus. Because of using digital sig¬
nals, much of the control circuitry operated only during the duration of the pulse.
In other words, the duty cycle for the circuit was considerably less as compared to
the continuously operating analog circuits. As a result, less heat was dissipated
and less energy was consumed by the circuits. Less power was needed, too, to
run these circuits, only because of the elimination of power-hungry filaments
associated with vacuum tubes.
Another major potential flaw associated with digital communications is that
it generates its own share of electrical noise. Since, mathematically, a square
wave or pulse is composed of an infinite number of sine waves, a square wave or
pulse can generate tremendous amounts of interference because of the infinite
number of frequencies being generated. Despite this problem, and because its
advantages far outweigh the disadvantages, digital control circuitry is becoming
extremely popular.
Section 11-8 Pulse Modulated Control Systems 359

11-8 PULSE MODULATED CONTROL SYSTEMS

There are several methods used for pulse modulation, all of which fall under the
heading of digital communications. The ones that we discuss here are the follow¬
ing:

1. pulse amplitude modulation,


2. pulse width modulation,
3. pulse position modulation, and
.
4 pulse code modulation.

These four methods are the most basic pulse modulation methods used in industry
at the present.

11-8.1 Pulse Amplitude Modulation (PAM)

Figure 11-10 shows a picture of a pulse at different periods of time demonstrating


how its amplitude varies with whatever data is being used to vary its height. The
pulse’s height is modulated by the input signal or data. This is called pulse ampli¬
tude modulation, or PAM. As an example, an input voltage of 3 volts would
produce a pulse height that would be twice as high as an input voltage of 1.5
volts. This is assuming that the pulse’s height varies linearly with the input. This
type of digital system requires a very quiet no-signal or background noise condi¬
tion in order for it to function properly. It is susceptible to the same noise prob¬
lems as an amplitude modulated system.

Figure 11-10 Pulse amplitude modulation.

11-8.2 Pulse Width Modulation (PWM)

Figure 11-11 represents a pulse whose duration or width is modulated by an input


signal. The amplitude of the pulse remains the same. This type of system is called
pulse width modulation, or PWM. The receiver for this type of transmission must
360 Chapter 11 System Communications

Figure 11-11 Pulse duration modula¬


tion.

have the capability of measuring the duration of the received pulse and translating
the measured amount into an equivalent voltage, current, or other data form. This
system has a better noise immunity than the PAM system. As a matter of fact,
there are similarities between the PAM system and FM as far as how each per¬
forms under noisy conditions.

11-8.3 Pulse Position Modulation (PPM)

This type of modulation is shown in Figure 11-12. Pulse position modulation is


somewhat more complicated than the other methods already mentioned. It con¬
sists of a pulse of fixed width and amplitude. However, its position constantly
shifts relative to a known time or position reference. The pulse’s position relative

Figure 11-12 Pulse position modulation.


Section 11-8 Pulse Modulated Control Systems 361

to this reference is varied by the input data and determines the magnitude of that
input. Its noise immunity is somewhat like that of a PM system, meaning that it is
somewhat better than PWM and certainly better than PAM.

11-8.4 Pulse Code Modulation (PCM)

Of the four listed methods of pulse communications, pulse code modulation is the
one modulation type most extensively used today. It also has the most varia¬
tions. One form of PCM is seen in Figure 11-13. In this type of pulse modulation,

Binary Number 1 0 1

■*- Time

Figure 11-13 The decimal number 10 represented by the 8-4-2-1 binary code.

a group of pulses are combined to form a coded group to represent an alphanu¬


meric byte. Out of the group of four pulses in Figure 11-13 (the third pulse from
the right is missing), the binary number 1011, or decimal 11, is formed. This
grouping is based on the so-called 8-4-2-1 binary system. Which pulses show up
in which four possible locations in the grouping determines the magnitude of the
decimal number. Additional groups of four binary bits can be formed to create
larger magnitude decimal equivalent numbers. Table 11-1 is the truth table for the
8-4-2-1 system. (A truth table is a listing of all possible logical states that a

TABLE 11-1 TRUTH TABLE


FOR THE 8-4-2-1 CODED
BINARY SYSTEM

Decimal no. 8 4 2 1

0 0 0 0 0
1 0 0 0 1
2 0 0 1 0
3 0 0 1 1
4 0 1 0 0
5 0 1 0 1
6 0 1 1 0
7 0 1 1 1
8 1 0 0 0
9 1 0 0 1
362 Chapter 11 System Communications

coded group of symbols can have for a possible set of conditions.) As an example,
if we wanted to represent the decimal number 359 using the 8-4-2-1 code, we
would write the binary equivalent as 0011 0101 1001. The numbers 8-4-2-1 are
the weights for this code. That is, noting which columns have Is occurring in
them, and then merely adding the weighted values of those columns, will produce
the equivalent decimal number value. Notice also in Table 11-1 that we used only
10 of the 16 possible states that exist using an 8-4-2-1 coding scheme. The
remaining 6 states are referred to as forbidden states, and should these binary
numbers crop up in a digital automatic control system, erroneous data has obvi¬
ously been generated. A digital system can be made to look for these errors to
prevent system errors.
There are numerous other binary coded systems that are used in industry for
digital control. Each has its own particular advantage for usage. Table 11-2 lists
some of the more common five-bit codes, and Table 11-3 lists some of the more-
than-five-bit codes being used.

TABLE 11-2 FIVE-BIT BINARY CODED TABLE 11-3 COMMON


DECIMAL (BCD) CODES MORE-THAN-FIVE-BIT CODES

Decimal 2-out-of-5 63210 Johnson code 51111 Decimal 50 43210 9876543210

0 00011 00110 00000 00000 0 01 00001 0000000001


1 00101 00011 00001 00001 1 01 00010 0000000010
2 00110 00101 00011 00011 2 01 00100 0000000100
3 01001 01001 00111 00111 3 01 01000 0000001000
4 01010 01010 01111 01111 4 01 10000 0000010000
5 01100 01100 mil 10000 5 10 00001 0000100000
6 10001 10001 11110 11000 6 10 00010 0001000000
7 10010 10010 11100 11100 7 10 00100 0010000000
8 10100 10100 11000 11110 8 10 01000 0100000000
9 11000 11000 10000 mil 9 10 10000 1000000000

The 2-out-of-5 code has a built-in parity check provision. Parity checking is
a method of counting all the Is and 0s to make certain there is the required total
number or required odd or even number of Is or 0s present for that particular
character. Parity checking is often used in automatic control systems to check
data validity. In the 2-out-of-5, there are always two Is. The same is true for the
8—6—4—2— 1 code. In the 5-1-1-1-1 code, the code is self-complementing. That
is, the complement of a decimal number has a binary complement for its binary
equivalent. For instance, the 9’s complement of decimal 2, from Table 11-2, is
decimal 7 (9 — 2 = 7). (The term 9’s means to subtract the decimal number from 9
to find that number’s 9’s complement.) The l’s complement of binary 00011 (deci-
Section 11-9 Frequency Shift Keying 363

mal 2) is binary 11100 (decimal 7). (The l’s complement of any binary number is
generated by merely changing all Os to Is and all Is to Os in that number.)
Therefore, the complement of the decimal using this coding scheme has a binary
l’s complement equivalent. Another example: The 9’s complement of decimal 4
is 5 (9 - 4 = 5). The complement of binary 01111 (decimal 4) is binary 10000
(decimal 5). Because of this self-complementing feature, error checking is made
easier and the electronics needed to decode the binary numbers is also made
easier.
Looking at another coding scheme, the 50 4321 code, we have another built-
in parity check system. The 01 and 10 bits are even parity checkers. That is,
when all the Is are added in the entire binary number including the parity bits
themselves, we always come out with an even number of Is. As a matter of fact,
we always come out with two Is in every number. This makes another fast and
accurate system error check.
Another popular form of PCM is the ASCII code. The letters ASCII (pro¬
nounced askey) stand for American Standard Code for Information Interchange.
This is the code that virtually all personal computers use today for communication
purposes between computer systems and their peripheral equipment. Figure
11-14 shows the truth table for the ASCII code. Notice that this coding scheme is
referred to as an eight-bit code, whereas only seven bits are explained in the truth
table. The reason for this is that the eighth bit is used as a parity check for the
other seven.

11-9 FREQUENCY SHIFT KEYING

Frequency shift keying (FSK) has been around for many years in the field of
communications. This system uses two audio tones that are keyed alternately off
and on. One tone represents a 1 while the other represents a 0. Teletype systems
have used this system extensively for transmitting messages over telephone lines
and over radio. FSK is probably used more today than ever before because of the
advent of the computer. Computer information is transmitted by this method over
both telephone and radio using a device that is called a modem. The word is a
composite of the two words modulator and demodulator. The tone frequency
pairs often used for this application are 1,270 Hz and 1,070 Hz, and 2,225 Hz and
2,025 Hz. Other tone combinations are also used as long as the receiver has been
designed to receive and separate those tones.
For radio transmission, the tones are fed directly into the transmitter’s mod¬
ulator audio input. The major disadvantage to this method is that all data must be
.transmitted serially. In other words, in order to transmit a “word” of informa¬
tion, the binary equivalent must be transmitted bit-by-bit, one at a time in succes¬
sion. This is called serial transmission. The problem with this method is that it is
very slow. However, data can be transmitted over relatively long distances using
this process. The other mode of transmission, called parallel transmission, allows
364 Chapter 11 System Communications

BIT 7: 0 0 0 0 1 1 1 1
Legend:
BIT 6: 0 0 1 1 1 0 0 1 1
NULL Nu I I/I die
BIT 5: 0 0 1 0 1 0 1
SOM Start of Message
EOA End of Address
BIT 4:

BIT 2:

BIT 1:
BIT 3:

EOM End of Message


EOT End of Transmission
WRU "Who Are You ?"
ik 1 k RU "Are You . . . ?"
0 0 0 0 NULL DC0 b 0 @ P
BELL Audible Signal
Format Effector
0 0 0 1 SOM DC, ! A Q
1 HT Horizontal Tabulation
— —

/ r
SK Skip (Punched Card)
0 0 1 0 EOA dc2 2 B R
LF Line Feed
Wab Vertical Tabulation
0 0 1 1 EOM dc3 # 3 C S FF Form Feed
1 1
CR Carriage Return
dc4 N
0 1 0 0 EOT $ 4 D T 50 Shift Out
(STOP) A
S 51 Shift In
0 1 0 1 WRU ERR % 5 E U Device Control Reserved
S DC0
lJ for Data Link Escape
N
0 1 1 0 RU SYNC & 6 F V G D©- DC3 Device Control
A N
/ DC4 (Stop) Device Control (Stop)
S '
E Error
0 1 1 1 BELL LEM 7 G w S ERR
(APOS) C)
| _ SYNC Synchronous Idle
G Logical End of Media
1 0 0 0 S0 ( 8 H X LEM
N Separator (Information)
S0-S7
HT E
1 0 0 1 ) 9 I Y b Word Separator (Space,
// SK Si [)
Normally Nonprinting)

1 0 1 0 LF S2
*
J z < Less Than

— > Greater Than

1 0 1 1 + K [
t Up Arrow (Exponentiation)
VjAB S3 /
't Left Arrow (Implies/
Replaced By)
1 1 0 0 FF S4
/
< L \ ACK
(COMMA) \ Reverse Slant
ACK Acknowledge
1 1 0 1 CR - = M ]
S5 © ©ESC Unassigned Control
Escape
1 1 1 0 SO S6 • > N t ESC
DEL Delete/ldle

1 1 1 1 SI / ? 0 X— DEL
S7
r

Example: Character "R" is represented by 0100101

Figure 11-14 ASCII eight-bit code truth table. (James Martin, TELECOMMUNICA¬
TIONS AND THE COMPUTER, 2nd ed., copyright 1976, p. 308. Reprinted by permission
of Prentice Hall, Inc., Englewood Cliffs, N.J.)

the user to transmit all bits of an entire “word” simultaneously. However, this
requires using many channels of transmission media. That is, one channel for
each bit is required. If wire is going to be used for the conducting media, then
multiconductor wire has to be used. We discuss this and other transmission me¬
dia in the next section.
Section 11-10 Types of Transmission Media 365

11-10 TYPES OF TRANSMISSION MEDIA

There are five major types of transmission methods and media that are presently
being used to transmit and receive control signals. These are:

1. wire conductors,
2. radio waves,
3. fiber optics,
4. light waves, and
5. pneumatic transmission lines.

As pointed out in the previous section, the faster method of digital data transmis¬
sion is parallel transmission. If that is going to be the selected mode of data
control transmission in an automatic control system design, then either multicon¬
ductor wire cable must be used or multistrand fiber optical cable. Otherwise,
serial transmission can be employed, using any of the four aforementioned meth¬
ods. Keep in mind, though, the much slower speed of transmission. Despite this
one limitation, serial transmission of PCM information has had widespread use in
automatic control applications.
Since multichannel transmissions for PCM systems are cumbersome and not
all that convenient to design (in many instances, you may have to convert analog
data to digital data and then convert digital data back to analog data, requiring A/
D and D/A circuits), much simpler systems avoiding digital circuitry altogether
may be used. The transmission of data by means of carrier transmission may be
used instead. An example of this was illustrated in Figure 11-1 where the three
phase control voltages were converted to AM signals (an FM or PM signal conver¬
sion would probably have been better to take advantage of the superior noise
immunity characteristics). You can see the problem that would be created if it
had been decided instead to convert each of the three phase voltages to say, eight-
bit binary data and then to transmit that data to the receiving site at the tower.
Each of the three phases would require a minimum of eight channels of transmis¬
sion, not to mention the A/D and D/A conversion circuitry needed. Furthermore,
to transmit parallel information over such long distances and to expect it to arrive
at the receiver all at the same time to maintain proper phase alignment, is ex¬
tremely difficult with the present state of the technology. With all of this in mind,
let’s take a look at our four basic transmission media.

11-10.1 Wire Conductors


;

Wire conductors used for instrumentation and control signal applications in indus¬
try are usually shielded and heavily insulated to withstand abrasion. A wide
variety of these conductors are on the market today. In most cases, an outside
shield surrounding the inner signal conductor is usually kept at ground potential to
366 Chapter 11 System Communications

prevent interference from occurring from outside stray unwanted signals. This
type of cable is often referred to as coaxial cable, or simply coax. Coax is manu¬
factured with different impedances, the two most common impedances being
approximately 50 ohms and 75 ohms. Of these two, the 50-ohm coax is the more
popular.
Recently, flat cable, or ribbon cable as it is more popularly called, has
become popular for control signal applications. The reason for this is due to the
popularity of the microcomputer. The computer has made it necessary to trans¬
mit and receive digital data using parallel transmission techniques over short
lengths. Ribbon cable can also be purchased with shielding if desired; this is
recommended especially for industrial environments.

11-10.2 Radio Waves

A common means of transmitting data is by radio transmission. In recent years,


some spectacular results have been obtained using this method in the science of
satellite technology. The automatic control of satellites and extraterrestrial ro¬
botic vehicles over immense distances has proven beyond all doubt that this is a
viable means of data transmission. In some cases, more than one frequency is
used in order to handle more than one channel of data as seen in Figure 11-15.
One frequency is the command frequency for transmitting the setpoint informa¬
tion, while the other is the response frequency. In some cases, elaborate schemes
are used where only one frequency would be used but several subcarriers would
be transmitted and received on that same frequency. The subcarriers would

Figure 11-15 Long-distance robotic remote control.


Section 11-10 Types of Transmission Media 367

either be audible frequencies or at lower radio frequencies below the carrier’s


frequency. The earth-based transmitting power needed for satellite or robotic
control is in the hundreds or thousands of watts. The onboard transmitting power
beaming back to earth would only be several or tens of watts due to power supply
energy and weight limitations.

11-10.3 Fiber Optics

Fiber optics represents the newest developed media for the transmission of data.
Fiber optics involves the using of glass-like fibers that act as optical light pipes. A
modulated light source, usually a laser, is used for transmitting light beams that
act as carriers through the fiber. At the other end of the fiber cable, a photocell
circuit is used to demodulate the light beam carrier to remove the information.
The advantage of fiber optics over radio or telephone transmissions is that little
cross-talk or interference between channels is experienced. Different colored la¬
sers can be used to create different channels in the same fiber. At the receiving
end, the receiver can filter out its particular color of data and exclude all others,
much like the audio tone filtering system described earlier. In addition, many
fibers can be used to handle many different channels. It’s not unusual to see one
cable carrying many hundreds of fiber channels.

11-10.4 Light Waves

The transmission of data by light waves is another relatively new technology. The
theory of operation has been around for a good many years, but it has been only
recently that practical systems have been worked out. Basically, the only differ¬
ence between this system and the one just described involving fiber optics is that
no fiber optics are used. Instead, the modulated light beam information is trans¬
mitted directly either through the atmosphere or through space to the photosensi¬
tive receiver. Again, lasers are used to generate the powerful narrow beams of
light necessary for this kind of system to work. The disadvantage with this system
is that it is susceptible to atmospheric and man-made interference. The fiber optic
system, on the other hand, is virtually interference-free. Fiber optics require a
considerable installation investment (cost of cable, installation labor, etc.)
whereas the light communications system has zero installation costs between the
transmitters and receivers. At the present time, there are very few installations
using this method of remote automatic control; however, as time goes on, this
system will be used more and more.

11-10.5 Pneumatic Transmission Lines

The method of transmitting control signals by using pneumatic lines is the only
practical nonelectrical method presently being used in industry. This is perhaps
one of the simplest methods also.
368 Chapter 11 System Communications

As the name implies, an air line, usually made of copper tubing or other
durable tubing material, is filled with pressurized air. The tubing acts as a conduit
to transmit control signals from a transmitting device, capable of modulating the
air’s pressure inside the tube, to a receiver at the opposite end of the tube, used
for the purpose of interpreting the modulated signals. The transmitter is generally
comprised of a “flapper” valve whose open and closed position is controlled by a
positioning lever or member. The position of this lever could be in direct response
to a change in a position, a temperature, pressure, or flowrate of a fluid. Figure
10-28 in Chapter 10 shows just such a system. The amount of bypassed air created
by the flapper causes a proportional decrease or increase of air pressure inside the
air tube. The receiver at the tube’s opposite end then interprets the rise and fall of
air pressure as a proportional change in the measurand at the other end. In the
simpler control systems, the pneumatic controls are set up strictly for an on-off or
bang-bang servo application rather than a proportional application.
There are two inherent drawbacks to a pneumatic control transmission sys¬
tem. For one, the maximum practical length of pneumatic transmission line can
be restrictive. Transmission line lengths in excess of 600 feet are rarely used.
Pressure signals become smoothed or very poorly defined beyond this length.
Also, pneumatic systems are limited by relatively low transmission rates because
of the air’s compressibility. Transmitted pressure fronts comprising the signal
information can travel no faster than the speed of sound, thereby limiting the
response time of the overall system. The major advantages of this kind of system,
in light of the overwhelming disadvantages just mentioned, are low cost, low
maintenance, and that the systems can operate in volatile atmospheres. Because
of these three reasons, pneumatic systems are still in popular use today.

SUMMARY

The study of control system communications is extremely important, with the


increasing sizes of these systems. One of the biggest problems confronting the
design of such systems is the speed of communications. Obviously, digital serial
communications is the only practical method that can be used for long-distance
data communications. On the other hand, digital parallel communications is fairly
fast but very limited in the distances it can travel. (This may change in the near
future, however, with the advent of multichanneled laser space transmissions.)
This is way carrier-type transmissions have found many applications in this area.
Basically, carrier transmissions were developed to transmit the analog informa¬
tion of the earlier automatic control systems and to do it reliably. Considerable
development is needed to make fiber optics cost-competitive with the wire con¬
ductor systems for long-distance digital control system applications. However, it
will be only a matter of time before this happens. Already, many wire conductor
communications systems are rapidly being replaced with fiber optics. Because of
References 369

the much wider bandwidth and multichannel capabilities of fiber optics, so much
more digital information will be able to be transmitted and received.

REVIEW QUESTIONS

11-1. List the three major categories of data transmissions and give a character¬
istic of each.
11-2. Describe two examples of bang-bang servo systems. What main feature
differentiates a bang-bang system from a proportional control system?
11-3. Why is FM used more frequently than AM for the transmission of analog
data signals?
11-4. Explain how a Schmitt trigger works in a noise filtering system.
11-5. Explain the difference between an FM data signal and a PM data signal.
Are there any advantages in using PM signals versus using FM signals in a
communications system?
11-6. What is a major concern in transmitting pulsed or square-wave-type sig¬
nals? Explain in detail making references to sine-wave and square-wave
theory.
11-7. List the four basic forms of pulse modulation and describe characteristics
of each.
11-8. Describe what is meant by the term self-complementing in reference to
certain binary-type codes.
11-9. Cite an advantage and disadvantage of serial data transmission; do the
same for parallel data transmission.
11-10. Cite an advantage for using pneumatic data transmission versus using
electrical transmission methods. Cite at least two disadvantages for using
pneumatic transmissions.

REFERENCES

American Radio Relay League, The Radio Amateurs Handbook, Newington,


Conn.: ARRL Inc., any recent annual edition.
Buchsbaum, Walter H., Practical Electronic Reference Data, Englewood
Cliffs, N.J.: Prentice-Hall, Inc., 1978.
Vergers, Charles A., Handbook of Electrical Noise: Measurement and Tech¬
nology, Bike Ridge Summit, Pa.: TAB Books, 1979.
—\12\ _

Computers and
Automatic Control
Systems

12-1 INTRODUCTION

In Chapter 11 we discussed methods of transmitting and receiving digital informa¬


tion being used for automatic control systems. These systems were obviously
using some form of computer or hardware digital devices to generate and handle
the many bits of data. Because of the impact that computers are having in indus¬
try and on automatic control systems in particular, it’s necessary to understand
how the digital computer works. In Chapter 12 we pick the computer apart, using
a typical microcomputer as our subject, to see how control data is processed.
With a basic understanding of how computers work, you won’t feel the intimida¬
tion that many feel when they first approach a computer or any other kind of
digital system. In order to work with digital automatic control systems, it’s neces¬
sary to understand how computers work and how they work with control systems.

12-2 WHAT IS A MICROPROCESSOR?

In order to understand what a microprocessor is, we must first understand what a


CPU is. The letters CPU stand for Central Processing Unit. Before the advent of
the microcomputer, the microprocessor didn’t exist. All computers, instead, had
CPUs. The CPU was the brain of the computer. Its purpose was to perform all of
the computer control, all the handling of input and output data transfers between
circuits, all the arithmetic and logical operations, and it performed all the instruc¬
tions obtained from the various memory banks within the computer.

370
Section 12-2 What is a Microprocessor? 371

As time went on, the first computers began evolving into more powerful and
compact machines. This was due to the integrated circuit chips becoming more
sophisticated and more dense as more circuits were crowded into their limited
containment areas. Medium-scale integrated circuit chips, or MSI chips, gave
way to large-scale integrated (LSI) chips, until finally, what was originally called a
CPU and was spread out over several square inches of circuit board became a
single LSI chip, dubbed the microprocessor chip.
A typical 40-pin microprocessor containing all the circuitry just mentioned,
along with its general size, is illustrated in Figure 12-1. The processor chip itself

i k

2.00 In.

-► Figure 12-1 A close-up look at a


0.75 In. microprocessor.

most likely covers an area no greater than the dashed-line area shown inside the
figure’s sketch of the IC’s housing itself. Because of the rapid development of
computer chips and the computers themselves, the standardization of the nomen¬
clature has been left behind in this flurry of activity to survive on its own. As a
result, what comprises the CPU and what comprises the microprocessor for a
microcomputer nowadays is a little confusing. As it turns out, they are one and
the same. The microprocessor chip is the microcomputer’s CPU. The circuit
board on which the microprocessor is installed is called the mother board. The
box or cabinet that contains the mother board and all the other peripheral devices
is called the computer’s CPU box (Figure 12-2).
Let’s look more closely now at the organizational structure of a micropro¬
cessor chip. Figure 12-3 shows a box diagram of a rather simplified version of a
typical microprocessor. It really isn’t necessary to understand how this chip
works to use a computer, but it does become necessary to have some idea as to
what goes on inside these chips if you become involved with circuit design or with
some aspect of trouble-shooting. Besides, it gives you a much better feeling for
what is going on inside computer-controlled circuitry such as digitally controlled
automatic control circuits.
The ALU (or arithmetic-logic unit) of a microprocessor performs all the
mathematical and logical chores on the data that is supplied to it. It has two
inputs: One input is from the accumulator, the other from the sequence control¬
ler. Generally, the ALU does only two forms of math; it adds binary numbers and
372 Chapter 12 Computers and Automatic Control Systems

Figure 12-2 Complete personal computer showing CPU. (Courtesy of HiTech


International, Inc., Milpitas, Calif.)

subtracts them. Technically, it only adds binary numbers. Subtractions are done
through what is called 2’s complement addition. Here is an example, except we
use 10’s complement addition (for reasons we explain in a moment): To subtract,
say, 4 from 7, we instead add the 10’s complement of 4 to 7. The 10’s complement
of 4 is found by taking 9-4 (not 10 - 4) which is 5, and then adding 1. In our
case, the 10’s complement of 4 is 6. Then, adding 6 to 7 we get 13. We then drop
the 1 in front of the 3 to get our answer. We used the 10’s complement only
Section 12-2 What is a Microprocessor? 373

CDNTRDL
INPUT

CLDCK
INPUT

Figure 12-3 Flowchart for a typical microprocessor.

because we were working with decimal numbers. In binary numbers, we would


use the 2’s complement. It would work like this:
7= 0111
Normal subtraction: — 4 = — 0100
3 = 0011

Using 2’s complement: The 2’s complement of 0100 is 1011 (the l’s complement or
reverse of 0100) + 1, which is 1011 + 1 = 1100. Then,

0111
+ 1100
10011
drop the “carry”

The answer is 0011, or decimal 3.


This routine seems a bit complicated and perhaps somewhat nonsensical at
first, but it is so easy to design digital logic circuits to perform the 2’s complement
for subtraction that it is used extensively. The binary numbers that are operated
on (called operands) by the ALU are supplied by the accumulator. The results of
an ALU arithmetic procedure are dumped back to the accumulator. Notice that
attached to the ALU is a carry register. Its purpose is to handle the carry routines
needed to properly manipulate the binary numbers, such as the foregoing routine
needed to discard the carry to obtain the final proper answer.
374 Chapter 12 Computers and Automatic Control Systems

The purpose of the accumulator is to accumulate data for the ALU and to
temporarily store completed data from the ALU. The accumulator’s instructions
come from the control bus that originates in the ALU. Essentially, the ALU tells
the accumulator when to operate and when not to operate.
The data register is another temporary holding area for data while an instruc¬
tion is being decoded. It also acts as an overflow for the accumulator. Usually,
the register can hold only one eight-bit data word. The data register then goes into
an instruction decoder to a sequence controller where the decoded instructions
coming from the ALU (these are usually math instructions such as x, + , etc.)
are cleared for operation and synchronized with other control input data coming
from outside the microprocessor. Notice that there is a clock input also going into
the sequence controller to do the actual synchronizing of all commands. The
clock insures that all data is in pace with their control signals and that everything
arrives at a particular location in the data processing at the correct time.
Let’s look at the program counter now. This is another memory area that
contains the address of the next command or instruction that is to be run in a
program. The program counter’s main job is to sequence the order of each pro¬
gramming instruction. The numbering sequence may be rearranged however, if
the programmer wishes to “jump around’’ within the program. This is done by
issuing branching instructions within the program’s routine. The program counter
is controlled by the aforementioned control bus. Its output goes into the address
register.
The address register is used to temporarily store the address of a data word’s
memory location. What is interesting about this feature is that this gives the
programmer the option to alter the data’s address, thereby changing the routine or
sequence of which data is to be processed. Again, this area of the CPU is con¬
trolled by the control bus. The output of the address register goes to the CPU’s
temporary memory, called random access memory (RAM), and to the I/O port¬
ing. The I/O porting refers to the circuitry that controls the inputting and output¬
ting of the floppy drives and hard disks. This porting also controls the sending of
data to the monitor and printing devices, not to mention receiving data from the
keyboard.
The microprocessor, or processor as it is often shortened to, “talks’’ to the
outside world by way of three different routes or paths. One path is called the
data bus, another is the address bus, and the third path is called the control bus.
It’s the data bus that distinguishes one microprocessor chip from another. The
data bus is a group of connections or pins on the IC chip that allow for the two-
way exchange of data or for the inputting of instructions to the processor. There
may be 4 lines, 8 lines, 16 lines, or 32 lines. This group of lines represent the
number of digits that can be made up to form a binary “word’’ that can be
manipulated by the microprocessor. That is, these lines represent the parallel
processing capacity of the processor itself. Therefore, these microprocessors are
referred to as 4-bit processors, 8-bit processors, 16-bit processors, etc. Table 12-1
lists some of the more popular microprocessors that have been used in the past
Section 12-3 The Computer and its Software 375

TABLE 12-1 POPULAR MICROPROCESSOR


CHIPS PRESENTLY BEING USED
IN MICROCOMPUTERS

Chip number True data bus width Manufacturer

Z80 8 Zilog
6502 8 MOS Tech.
8086 8 Intel
8088 8 Intel
80286 16 Intel
80386 32 Intel
6800 8 Motorola
68000 16 Motorola

and are presently being used in microcomputers today, along with the names of
their manufacturers.
In Table 12-1, the reason for the term true data bus length is because some
computer designers can take, say, an 8088 chip and simulate a 16-bit processor
through manipulation of the data bus configuration, but in reality it is still only an
8-bit processor chip. The chips that have been listed here are those that comprise
the greater part of the microcomputer market. The term microcomputer is really
what is now more commonly referred to as a personal computer or PC. Many
PCs, as a result of containing large amounts of RAM (Random Access Memory)
and having several 10’s of megabytes of hard-disk memory available, are ap¬
proaching the sizes of minicomputers.

12-3 THE COMPUTER AND ITS SOFTWARE

Let’s look at a personal computer a little more closely, especially at the mother
board contained by its CPU box, as illustrated in Figure 12-4. On this mother
board, the CPU (which, by the way, is an 80286 16-bit processor) is installed, with
a heat sink, in the upper center on the board, alongside the hard drive enclosure.
Slightly above and to its left we see the 80287, a 40-pin chip. This is the micropro¬
cessor’s matching math coprocessor. A math coprocessor is an auxiliary chip that
speeds up the math computation speed of the computer. It’s been especially
designed to perform high-speed calculations and other related routines. If you’re
planning to use a computer that’s going to be used for math-intensive applications
for automatic control applications, or for any other applications, you should con¬
sider installing one of these. Just be sure that it is the matching coprocessor for
your particular microprocessor and that it’s designed to work with your micropro¬
cessor’s operating speed.
Near the lower right-hand corner of the mother board we see several rows of
ICs which are the computer’s RAM chips. RAM is the computer’s temporary
376 Chapter 12 Computers and Automatic Control Systems

LITHIUM BATTERY 80287 HOUSING FOR HARD DISK


COPROCESSOR \
CHIP 80286 UP CHIP \

EXPANSION SLOTS (8) BIOS ROM CHIPS RAM CHIPS

Figure 12-4 A mother board.

memory for holding data that is being worked on. The memory is temporary, in
that when the computer is turned off, the memory information is lost forever.
However, when energized, information can be read from this memory and also
“written” into this memory. This is an important point to realize, because not all
memory has this read/write capability. For instance, the two chips seen with the
foil labels at the lower edge and left-of-center of the mother board are ROM (Read
Only Memory) chips. These chips have data information that has been perma¬
nently implanted into the chips and can be read at any time. However, you can’t
write to these chips. These particular chips have a unique function in the com¬
puter. They comprise what is referred to as the BIOS for the computer. This is an
acronym which means Basic Input!Output System. In other words, these chips
contain operating commands that manage the operation of the computer’s several
input/output ports, which are attached to peripheral equipment such as floppy
disk drives, hard disk drives, back-up tape drives, etc.
In the left-hand portion of our mother board, we see several empty card
slots. These are the slots that accept additional printed circuit boards that are
dedicated to the operation of our computer’s peripheral devices. One slot may
contain a video control board that allows the operation of a monitor. Another
Section 12-3 The Computer and its Software 2>11

board may be the controller board for the floppy disk drives or the hard disk.
Another board may be a board for allowing the computer to “talk” to another
computer over the telephone. (This board is called a modem.) It’s important to
realize the differences between these two kinds of memory devices. A floppy disk
drive is an electromechanical device that allows the computer operator to read
and write data information to a disk for permanent storage applications. They are
quite slow in operation as compared to a RAM chip. However, relatively large
amounts of data can be stored and accessed in this way. Typical capacities range
from 360 kilobytes to as much as 1.4 megabytes. A hard disk is also an electrome¬
chanical device, but is much faster in operation as compared to a floppy disk
drive, and can hold tremendous amounts of data. Typical storage capacities range
from 10 megabytes to several hundreds of megabytes. Hard disks are still consid¬
ered to be slow compared to RAM storage, but at least the stored information is
nonvolatile, like the floppy. That is, the memory is not affected by loss of power
to the memory device as is the case with RAM chips. (The hard disk, on the other
hand, is susceptible to stray magnetic fields much like any other recording mag¬
netic media, and also to dust or dirt particles that may be present between the
media and the reading and writing heads of the hard disk mechanism.) The RAM
chip is considered to be a volatile memory device. A nonvolatile RAM chip
memory system is made possible by using batteries to hold up the memory. This
is a fairly common practice to prevent memory loss during power outage condi¬
tions.
Now that we have identified the major parts of a microcomputer, let’s look
at the different forms in which data may be written and programmed so that it can
be stored and processed by a computer. These different forms are the computer’s
software languages. Essentially, there are only three major categories of pro¬
gramming languages:

1. binary or machine code language,


2. assembly language, and
3. the higher level, or structured languages.

The most basic of software languages is machine language or code. This is a


language form comprised of nothing but binary bits or numbers. This is an ex¬
tremely tedious and lengthy method of writing and is not often used anymore.
Machine language is referred to as a low-level language. As a matter of fact, it’s
the lowest possible level one can use in writing software, since it is the language of
the computer. All data manipulations and all control signals are handled in binary
form on the computer’s mother board and peripheral boards. Sometimes when
)

using machine language, rather than enter binary numbers, often what was done
was to enter this binary information in hexadecimal form. This was done to
reduce the number of keystrokes and to reduce keystroke errors. The keyboard
378 Chapter 12 Computers and Automatic Control Systems

was in hexadecimal form, and some form of light-emitting diode (LED) display
was used on the earlier computers to keep track of programming progress.
The next lowest level software language is assembly language or code. This
is a language form comprised of mnemonic statements; that is, statements that
could be pronounced by the user for easy memorization. Each mnemonic state¬
ment represents a binary coded statement or several binary words that the user
would never see. However, should that mnemonic statement be used, a series of
binary codes would be generated. An assembler is used to actually perform the
conversion of the assembly language into actual binary words. In order to pro¬
gram in assembly language, the user must have a thorough understanding of the
computer’s hardware and architecture.
The category of programming languages covers a wide territory. The pro¬
grammer has a wide selection of programs to choose from. Higher level program¬
ming languages encompass those languages that have the user enter certain En¬
glish words of instruction that perform a particular action or manipulation of the
data. Perhaps one of the better examples of such a language is BASIC. By enter¬
ing the proper sequence of these commands or instructional words, the program
will do what the programmer has instructed it to do. The programmer is unaware
that for each of his or her instructions in the program, that instruction has been
automatically converted to machine language for the computer to run on. That
portion of the high-level program that does the high-level language-to-binary con¬
version is called an interpreter, or compiler. The major difference between an
interpreter and compiler is found in the methods used in generating the machine
code. A compiler reads the entire high-level program first and then generates the
machine code; the interpreter, on the other hand, reads one high-level instruction
or statement at a time, producing the machine code as it goes along.
Some high-level programs are higher level than others. For instance, there
are several statements in BASIC that occur frequently together to perform a
particular function. Another programming language will take those statements
and combine them into one statement so that all one has to do is type that one
word instead of having to type the several words as had to be done in BASIC.
Much time in programming can be saved using a higher level language. However,
programming versatility is lost when using such a language. This used to be one of
the major drawbacks; however, this is changing with more sophisticated programs
now being marketed. This allows the user to have greater flexibility in program¬
ming and still maintain the greater programming speed characteristics of higher
level languages. It’s interesting to note that some programming languages are
approaching using common everyday English expressions for the programming
commands. As a matter of fact, more than one common expression will perform
the same command. This means that the programmer, without having to resort to
a memorized command set of words as is now done, can instead issue a command
word using his or her intuition. The programming language will then refer to a
dictionary of terms that will interpret equivalent words and issue what it interprets
to be the desired command.
Section 12-4 An Example of Automatic Controlling Using a Computer 379

12-4 AN EXAMPLE OF AUTOMATIC CONTROLLING USING


A COMPUTER

Now that we have some understanding of how computers work and what software
is all about, our next task is to understand how all of this information is applied to
a computerized digital automatic control system. To begin with, we have already
gained some insight on how the controlling and analyzing of data is done. But we
back away now and take a look at the computer itself and at the surrounding
hardware that it can send information to and receive information from.
Let’s begin with analyzing a computer-controlled automatic control system
used for process control. As an example, the process we discuss here will be the
cold-chamber aluminum die casting process we discussed earlier in Section 10-9.
We wish not only to monitor the various important parameters that go into the
making of a good aluminum die casting part, but we also wish to control these
parameters and to make any necessary corrections, automatically. In other
words, we want to make an adaptive control system for our die casting machine.
Each process control system has its own particular set of parameters it must
monitor and control in order to insure producing a good part. For the die casting
process, these parameters are (refer to Figure 12-5):

1. metal temperature;
2. shot velocity (this is the velocity of the hydraulic ram or plunger used for
injecting the molten aluminum into the die);

Figure 12-5 Parts of a die cast machine. (Wayne AlofslJames R. Carstens, MECHANI¬
CAL MAINTENANCE AND EVALUATION OF DIE CASTING MACHINES, copyright
1987. Reprinted with permission by the Society of Die Casting Engineers, River Grove, III.)
380 Chapter 12 Computers and Automatic Control Systems

3. lubrication of shot tip (this has to do with the molten metal injection portion
of the cold chamber and is critical for keeping wear to a minimum between
the shot plunger and the shot sleeve itself);
.
4 location of ladle used for pouring metal into cold chamber shot hole;
5. dimensions of finished part;
6. proper part ejection;
7. tie bar tension on the die casting machine to maintain proper closure tonnage
on die during metal injection and cooling of part;
8. die temperature;
9. lubrication of the die casting machine’s toggle system (This is the mechani¬
cal portion of machine that mechanically amplifies the hydraulic pressure
used in creating the proper closing pressures on the die); and
10. electrical energy consumption of the die casting machine.

It takes at least one or perhaps two minutes to produce one part out of a die
casting machine. In other words, this is the cycle time for the process. Within
this cycle time, all 10 of the foregoing parameters must be monitored and any
corrections to the system made. In order to monitor and control a complicated
operation such as this using a computer, transducers must be installed in appropri¬
ate areas surrounding the process. Ideally, these transducers should have the
capability of producing digital signals. This would eliminate any analog-to-digital
conversions that would otherwise be needed for interfacing to the computer. In
addition, you would have the maximum circuit noise immunity needed for a safe
and reliable operating system.
In developing the software needed to supervise this system, a program will
have to be written that will scan each of our 10 listed areas. The outputs of our
transducers will have to be wired so that each output can be accessed at the
computer’s input/output ports. This is done so that each signal can be compared
to a prerecorded signal value that has already been considered to be an acceptable
value. If any of the transducers’ outputs are considered to be unacceptable, ap¬
propriate corrections will have to be made automatically.
The best way to develop and understand an automatic control and monitor¬
ing system is to make a flowchart of block diagram of the system. Figure 12-6
illustrates such a diagram for our die casting system. Block diagrams tend to be
uninteresting, but they do contain a wealth of information. Be sure to study
Figure 12-6 closely and make sure you understand each line and notation. If you
do this, you’ll have little problem understanding any other systems you may have
to work with in the future.
Figure 12-6 shows what appears to be a mechanical scanning system where a
switch travels from one block system to another, making an electrical contact.
This is shown for illustrative purposes only. In reality, all electrical switching is
done electronically with no moving parts. As each portion of the die casting
machine is scanned, the computer sends back corrective information in the form
Section 12-4 An Example of Automatic Controlling Using a Computer 381

□UTPUT INPUT

Figure 12-6 A computerized control system for die casting.

of an error signal to that same area. As a matter of fact, both the scanning and
error corrections may be done several times a second, with the exception of the
following: shot velocity, finished part dimensions, ladle location, tie bar load, and
part ejection; these particular items can be checked only during or immediately
following the making of the part.
To illustrate the preceding remarks, let’s assume that our software has told
the input/output (I/O) port on the computer to read the metal temperature in the
ladle as it is being scooped from the nearby furnace. Thermocouples in the ladle
read the metal’s temperature and the resultant generated voltage. This voltage
382 Chapter 12 Computers and Automatic Control Systems

signal is sent to an A/D converter circuit. The circuit, in turn, sends a digital
signal to the computer, where its newly converted binary equivalent value is
compared to a desired temperature (probably around 1,250°F to 1,300°F). Any
difference in temperature is immediately sent back in the form of an error signal to
the furnace’s temperature controller, where any necessary adjustments are
made. Because of the massiveness of the furnace and the contained aluminum
charge, the response time, and consequently its settling time to the set point, is
going to be very slow, measurable in minutes.
The die casting machine’s hydraulically controlled shot producing system is
probably one of the more complicated and difficult systems to control automati¬
cally on the machine. Its description was given in Section 10-9. Because of the
speed with which a shot is made, it is difficult to make speed corrections with
generated error signals during the shot. It is much easier to make corrections on
the next shot. This is a typical problem for many mechanical devices having large
masses. Their response times are very slow compared to their total travel time,
and consequently, they must settle for next-cycle correction only.
The next item, shot tip lubrication, involves having to apply lubrication to
the tip of the ramming device that pushes the molten metal into the die. If this part
is not lubricated, excessive drag and wear is created in the shot sleeve. This is the
chamber in which the metal is poured. If the computer software detects any
slowing down of the ram during successive shots, this could be an indication of
lack of shot tip lubrication. A normal closed-loop system would create an increas¬
ingly larger error signal as the lubrication decreases, causing the velocity to be¬
come compensated with increased speed. Unfortunately, this would cause the
entire shot system to eventually destroy itself if no corrective steps were taken to
shut down the control system. However, with the adaptive control system that
we have described up to this point, additional lubrication would be automatically
added to the shot tip area to prevent its self-destruction from taking place.
Ladle location must be determined in order for the ladler or robot to accu¬
rately hit the pour hole in the shot sleeve as it pours the molten metal. This means
that the ladle’s position must be continuously monitored and corrected using the
methods outlined in Section 10-2.
The checking of finished part dimensions is performed by a vision system
similar to the one described in Section 10-3. This particular system is a go-no-go
system. If the part is within specifications, it is saved; if it isn't, the part is
recycled.
The part ejection system checks to make sure that the mechanical ejection
system built into the die has in fact ejected the part out of the die. This check may
be made by a displacement transducer built into the robot’s gripper that is used to
remove the part from the die. If the part is properly grasped by the gripper, the
transducer will transmit the proper displacement of the gripper. Otherwise, any
other signal value coming from the gripper will indicate a problem with the ejec¬
tion procedure. This system is basically another go-no-go type system rather than
a proportional feedback system.
Section 12-4 An Example of Automatic Controlling Using a Computer 383

The tie bar loading on the die casting machine is monitored and controlled
according to the description given in Section 10-9.
For the controlling of die temperature, the die contains temperature condi¬
tioning coils that have water circulating through them. Temperature mixing
valves are used to mix heated and cooled water to obtain the desired die tempera¬
tures. Because of the massiveness of the steel dies used, the temperature re¬
sponse time is quite slow, on the order of several minutes typically. This implies
that the settling times for any temperature corrections will be quite long. Installed
on the mixing valves are servovalves that adjust the valve stem positioning. The
die’s temperature is monitored by a scanning infrared detector that measures
temperatures over its entire face. These temperatures are then compared to the
desired temperature values stored in software. The software then calculates any
arithmetic differences in these values and sends that difference back to the die’s
water mixing facilities in the form of an error correction to operate the appropriate
valves for temperature correction.
General machine lubrication may be checked by sonic transducers (micro¬
phones) placed at the important bearing points on the machine. The sound output
of these detectors can then be compared to acceptable output levels. If any levels
are exceeded due to plugged or broken lubrication lines, alarms may be sounded
or lights flashed to get the attention of the machine supervisor. This system is
another go-no-no control system.
Machine energy consumption can be checked by comparing the electrical
current draw of the machine and comparing that information to what is considered
a normal current draw. The current may be monitored by a digital ammeter
circuit. Any increase in current draw may be an indication of binding members on
the machine due to member failure or a failed lubrication system.
As was pointed out during the describing of each of the machine operating
parameters, some of the parameters listed in Figure 12-6 obviously don’t require
an analog automatic feedback control. The finished part dimensioning, part ejec¬
tion, shot tip lubrication, general machine lubrication, and machine energy con¬
sumption areas are all go-no-go decision-making operations. Instead, these may
be thought of as bang-bang automated systems. As in any control system, how¬
ever, we want our computer control system to coordinate all of the various opera¬
tions of our die casting process; this means writing a rather extensive software
program to do all of this.
Again, we use a flowchart to see how we might write the software program
to automatically control our die casting operation. This is a very generic pro¬
gram. That is, it is merely to be an overview for a far more detailed program. The
choice of programming language used for this project is a choice left up to the
programmer and the computer that will be used to run the program. Regardless of
the language used, the programming logic remains the same. Figure 12-7 shows a
flowchart that could be used in aiding the programmer in writing the software for
controlling our die casting process. It must be pointed out that this is a very crude
chart. No safety interlocks are shown, and many other auxiliary control circuits
384 Chapter 12 Computers and Automatic Control Systems

START DF MACHINE
CYCLE

Figure 12-7 Flowchart for programming the automatic control of a work cell of a die cast
machine operation.

are also missing. Also, there are many other ways to construct the programming
flow to create the same end results.
Looking at Figure 12-7, the flowchart begins with checking the furnace tem¬
perature to make certain that the aluminum which is to be used in the die casting
machine is at the proper pouring temperature. A sensor on the furnace supplies
the temperature data and is compared to the set point temperature. If an error
signal results, the machine’s operation is stopped so that the furnace’s tempera-
Section 12-4 An Example of Automatic Controlling Using a Computer 385

DIE
TEMPERATURE CHECK DIE
SENSOR TEMPERATURE
CHECK
TEMPERATURE
AND ADJUST

NO INHIBIT MACHINE
SET POINT
(ERROR)
YES
(ERROR=0>
CURRENT CHECK MACHINE
SENSOR CURRENT DRAW
CHECK
CURRENT DRAV
AND ADJUST

NO
SET POINT INHIBIT MACHINE
(ERROR)
YES

CLOSE MACHINE
DN DIE

TONNAGE CHECK TIE BAR


SENSDR TONNAGE

ADJUST
NO
SET POINT INHIBIT MACHINE TIE BAR
(ERROR)
YES TONNAGE
(ERROR=0)

START LADLE

POUR
POSITION CHECK POUR
SENSOR LOCATION

CHECK
NO LOCATION
SET POINT
(ERROR AND ADJUST
YES
(ERROR=0)

Figure 12-7 Cont.

ture can be readjusted. During this readjustment period, the temperature sensor
continues taking readings until the set point is reached, or until the error signal is
zero. When this occurs, the next parameter is analyzed, which is check shot
;

speed. Again, a sensor’s reading (from the previous shot) is compared to the set
point. If the shot speed sensor detected a greater velocity from the shot just
completed as compared to the set point velocity, the shot controller is told to
reduce the shot velocity for the next shot by an amount equivalent to the error
386 Chapter 12 Computers and Automatic Control Systems

REPEAT SHOT
CYCLE

NOTE* The bracketed sensor/set paint designation are the pair of signals
used to nake a yes/no decision.

Figure 12-7 Cont.

value. If the shot speed sensor detected a smaller velocity as compared to the set
point velocity, the shot controller is told to increase the next shot velocity an
amount equivalent to the error value. However, in addition to this, the tip lubrica¬
tion circuit is also checked to make sure proper lubrication is present. This addi¬
tional check is made because a decreased shot velocity could be a result of lack of
tip lubrication. Following these checks, the software proceeds to check machine
lube.
As we continue down the flowchart, we encounter a series of functions
beginning with the pour command and continuing down through the start timer
and end timer commands. These two commands control a timer that determines
how long the die must be kept closed in order to allow the part to completely
freeze before reopening the die and ejecting the part.
In each of the parameter check categories listed in the flowchart, notice that
each parameter is confronted with a YES/NO routine. This is a common pro¬
gramming routine where checks have to be made on circuit conditions. These
Section 12-5 Using the Stepper Motor in an Automatic Control System 387

checks are logical functions carried out by software routine statements that ask
questions similar to, “If x-axis ladle position is <0.010 inch, then go to next
statement (i.e., YES), else, (i.e., NO) correct x-axis position to reduce error and
recheck.” All the statements in the flowchart to the right of each parameter check
are loop statements. These statements perform a correction and then cause a loop
to occur so that the parameter can be rechecked for error correction before
proceeding to the next parameter.
The foregoing process control system is admittedly rather detailed and spe¬
cific. It wasn’t intended to be a dissertation on the die casting process. However,
the analysis of the work-cell software routines is quite representative of many
automatic process control systems encountered in industry and is definitely worth
studying.

12-5 USING THE STEPPER MOTOR IN AN AUTOMATIC


CONTROL SYSTEM

The stepper motor is an ideal actuator for interfacing with the computer for
control. Since the motor itself responds to digital-like signals, it’s a natural device
for accepting commands and converting them into rotary motion. The theory of
operation of this motor was discussed in Section 4-17. What we discuss here is
how to control this motor with a computer and how the motor is used in an
automatic control application.
A typical stepper motor control system appears in Figure 12-8. This is an
entirely digital control system. A positional transducer is attached to the stepper
motor’s load to create the necessary error signal back at the comparator. The
directional control logic circuitry in Figure 12-8 is circuitry that converts the
mathematical sign of the error signal into the logic needed that will rotate the
stepper either clockwise or counterclockwise. (Be sure to review Section 4-17 if
you have forgotten how a stepper’s rotational direction is changed.) The software

Stepper

Set Point

Note: Dashed Lines Indicate a


Mechanical Coupling

Figure 12-8 An automatic control system using a stepper motor.


388 Chapter 12 Computers and Automatic Control Systems

needed to interpret the stepper’s position is laid out in the form of a flowchart in
Figure 12-9. This program is a continuous looping operation, since it must catch
any changing of the set point at any time made by the operator of the system. The
number of updates of the stepper’s position is controlled by the frequency of the
loops made in the software. This, in turn, is determined by the clock rate of the
computer’s CPU.

Loop

Note: The Number of Loops Per Second Made by the


Figure 12-9 Flow-charting the soft¬
Software Will be Determined by the Computer's ware for a stepper motor positional
Clock Speed. system.

A typical application of a system using a stepper motor for positional control


would be in the designing of a graphics pen plotter (Figure 12-10). In this system,
the plotter's pen location will be continuously changing, while at the same time in
the background the pen’s position would be continuously monitored and updated
by the algorithm seen in Figure 12-9. More than likely, this program would be
found in a ROM chip where it would then be protected from accidental alterations.
Section 12-6 Using a Programmable Controller for Automatic Control 389

Pulley Drive Pulley Drive

Carriage
Ways

Pulley

Pulley Optical Stepper


Drive Encoder Motor

Figure 12-10 A stepper motor pen plotter.

12-6 USING A PROGRAMMABLE CONTROLLER


FOR AUTOMATIC CONTROL

A programmable controller (or PC as they are frequently called) is a device that in


many respects resembles a personal computer (also called a PC). The first pro¬
grammable controller was produced in the middle 1960s for the purpose of replac¬
ing electrical relays. Up to that time, the electromechanical relay was used exten¬
sively to control industrial process control systems. All logical functions (the
AND, OR, NOT, NOR, etc. statements) were done in this manner. However,
the PC was a solid state device that used transistors, and later integrated circuits,
to perform all switching logic. A modern industrial PC is illustrated in Figure
12-11. In addition to using solid state switching for relays, the PC is programma¬
ble. The programming language most commonly used closely resembles the relay
language used by electricians and control engineers to draw out the control cir-
390 Chapter 12 Computers and Automatic Control Systems

INCOMING LINE WIRING TERMINALS


WIRING TERMINALS FOR 10 INPUTS

AU^*fiRAt>U?Y ft
cO*t«qu£s

BATTERY
COMPARTMENT
AUTO-MANUAL /
RWITCH EEPROM MEMORY
MODULE COMPARTMENT EXPANSION UNIT
CONNECTION

INPUT STATUS INDICATORS

WIRING TERMINALS DIAGNOSTIC INDICATORS


FOR 6 OUTPUTS
OUTPUT STATUS
INDICATORS

Figure 12-11 A programmable controller. (Courtesy of Allen-Bradley Co., Milwaukee,


Wis.)

cuitry schematics, called ladder diagrams (Figure 12-12). These diagrams are
comprised of all electrical relay symbols, and because of the familiarity that
electricians have with this relay logic, the relay ladder diagram programming
method with the PC has been very popular. As a matter of fact, to this day most
of the PCs being manufactured still use this form of programming. There are at
least two other forms of languages used, however, with the PC. These are Bool¬
ean algebra-based languages and mnemonic-based languages (similar to the com¬
puter assembly languages discussed earlier). Boolean algebra is a form of
algebraic logic that uses special math symbols in which the logic AND statement
is represented by a dot (•), the logic OR statement is represented by a plus ( + ),
and equivalent statements are represented by the equal sign (=). This type of
programming is more electronic in its approach as compared to the more conven¬
tional electrical approach.
The major advantages in using a PC is that, like a microcomputer, it is
programmable and it is also equipped to respond to the input signals coming from
switches and transducers (on the more elaborate models). It is also built to with¬
stand the hostile environment of an industrial installation. Most computers can't
survive in such areas. Unfortunately, most PCs lack the versatility of a micro-
Section 12-6 Using a Programmable Controller for Automatic Control 391

24 VAC

3, i push-button (pain) switch


4, °0£ 1 open contact on a Unit switch
5, _0_ ' holding or activating coil on a solenoid

Figure 12-12 Typical relay logic or “ladder diagram.”

computer as far as being able to process data, handle large memory requirements,
and run complex programs using the usual programming languages. Also, a mi¬
crocomputer allows the user to monitor the information being processed by using
the appropriate software. This is especially important when interpreting data
coming from a transducer. The major components of a PC are:

1. microprocessor,
2. nonvolatile memory,
3. input/output interfacing circuitry, and
4. programmer.

j
DEFINITIONS

MICROPROCESSOR The microprocessor has the same function as the


one found in microcomputers, except it usually doesn’t have the processing capa¬
bilities as the ones found in computers. The reason for this is to keep the PC
392 Chapter 12 Computers and Automatic Control Systems

manufacturing costs down. (This is changing, however. Some PC manufacturers


are beginning to install actual microcomputer chips to take advantage of the chip’s
additional computing power.)
NONVOLATILE MEMORY The nonvolatile memory has stored in it all
the commands and logic information needed to control the process. In case of a
power failure, the program wouldn’t be lost, since the memory would be “held
up” with batteries. This can become a serious problem with volatile memory
when power is restored and the controlled process has no homing instructions to
begin new cycling. Damage to equipment could result, and personnel safety could
be jeopardized.
INPUT! OUTPUT INTERFACING CIRCUITRY This is the circuitry that
allows the PC to control and to gather information from the process. The input
circuits and the output circuits are separated from each other through a modular
construction so that they can be segregated and moved around. Each I/O module
may handle over a dozen different circuits and is designed to handle AC, DC,
pulse, and BCD data in the more elaborate models.
PROGRAMMER The programmer is a system that allows the user to
program the PC. It may consist of a CRT (cathode ray tube) for following the
programming procedure, and a keyboard. The keyboard may be either attached
to the PC through a detachable cable, or the programming may be done through a
hand-held keyboard. Some hand-held models may even have a small liquid crys¬
tal display screen built into them for programming interpretation.
There are also auxiliary devices that can be used in conjunction with the PC
to allow the user to monitor process information. Figure 12-13 shows a device
designed to do this very thing. It’s called an Industrial Graphics Terminal. It has
the capacity to be programmable much like a computer so that it can display high-
resolution graphics. The graphics give a visual representation of the process be¬
ing monitored. It is interesting to note that some of the later models of PCs are
beginning to look and behave more and more like computers. They can handle
large numbers of input/output signals, perform complex math, create graphics,
and automatically control large process systems. In addition to all of this, they
are programmable in BASIC-like languages. One such system is called the
PCAM, or Programmable Controller and Monitor. It is specifically designed to be
used in automatic control systems where it can both monitor and control a pro¬
cess. A complete system hook-up is shown in Figure 12-14. Notice that one
PCAM unit is assigned to a particular work cell area or machine. A personal
computer is then used to gather data from each PCAM through a networking
system. This means that a supervisor can be in a remote location such as an office
with his or her computer and have the capability to control and monitor each of
these areas. Several dozen work cells can presently be controlled in this man¬
ner. As an added side note, the user can also gather management and diagnostic
data from those areas with this system.
Section 12-6 Using a Programmable Controller for Automatic Control 393

Figure 12-13 Graphics terminal for monitoring process data. (Courtesy of


Emory-Anderson, Inc., Comstock Park, Mich.)

Because this type of controller is programmable in a language similar to


BASIC, it is possible to write fairly complex programs. In other words, it is
possible to write a program similar to, or as extensive as, the one described by the
flowchart in Figure 12-7. This type of controller is especially designed to work
with transducers and control devices that are commonly used in automatic control
system applications. As an example, referring back to our flowchart of the die
casting operation, an ordinary PC could not handle and interpret the information
necessary for the adaptive controlling of the shot-making process. A PC AM de¬
vice, on the other hand, could not only interpret this information easily and make
the necessary adjustments automatically, it could also generate any necessary
monitoring information on a high-resolution graphics screen built into the unit.
394 Chapter 12 Computers and Automatic Control Systems

ETC.

Figure 12-14 A PCAM network. (Courtesy of Emory-Anderson, Inc. Comstock, Park,


Mich.)

Factory automation is becoming more and more prevalent these days. As a


result, the automation of processes has become increasingly important not only
from an economic standpoint but also from a safety standpoint. One of the great
concerns for a manufacturing process firm is how to remove its workers from
dangerous work situations. Automatic control of these processes is the answer.

12-7 METHODS OF COMPUTER INTERFACING

In working with electronic circuits, we learned that in order to hook up the output
of one circuit to the input of another, it was necessary to match their impedances
for maximum power transfer. Also, we had to be careful that proper voltage
protection existed so that we weren’t feeding, say, a 440 VAC three-phase circuit
into a 12-volt CMOS circuit. Obviously, we would have an instant problem devel¬
oping here. In hooking up circuits to a computer, we have much the same prob¬
lem. We are dealing with sensitive high-density solid state components inside the
computer which lack the ruggedness of the old vacuum tubes. We have to be
especially careful when designing control circuitry and wiring these circuits to
computers. Fortunately, the computer industry has standardized some of the
Section 12-7 Methods of Computer Interfacing 395

hook-up circuitry to reduce some of the difficulties just mentioned. Some of the
standardization procedures are still being formulated, but for the most part,
enough has been done to eliminate incompatibility problems between circuits and
computers.
Every industry has created its share of buzzwords for the English language,
and the computer industry is certainly no exception. When hooking up a circuit to
a computer, the term interfacing is used instead of the term hooking up. As a
matter of fact, the circuit being interfaced just may be a peripheral device for the
computer. In other words, it may be a circuit which assists the computer in
performing a particular function. What we are interested in here are the methods
used for computer interfacing. We have to be able to select the proper connec¬
tors, the cabling, and be able to understand the circuit enough to be assured that
the signals we want to swap back and forth between the computer and circuit are
digestible by each other.
Assuming that we are going to interface a digital automatic control circuit to
a computer so that the computer can monitor and control the circuit, we have to
ask ourselves, is the data being sent to the computer serial-type or parallel-type
data? This will determine the standard interfacing method to be used. If the data
coming from the circuit is serial, we would most likely use the RS-232C Inter¬
face. If the data is parallel, we would probably want to use the IEEE-488 Inter¬
face. An explanation of each of these along with others follows.

12-7.1 The 20 mA and 60 mA Current Loops

This is the oldest of existing methods used for interfacing electronic and electrical
equipment together. It was originally used for open-loop control of electrome¬
chanical equipment in communications (such as teletypewriters) and industrial
data processing. Distances between the control transmitter and the controlled
circuit at the receiving site can be as much as 2,500 feet. However, at these
distances considerable voltage-drop losses could result. These systems were used
for transmitting and receiving serial data. A notable disadvantage with this sys¬
tem is the very low rate of data transmission. This was due primarily to the
electromechanical switching devices that were used at the time. The highest rate
attainable was approximately 150 baud. (The term baud refers to the rate of data
transmission in units of bits/sec.) Two systems were popular; one involved using
current flows of 20 mA, while the other used 60 mA (Figure 12-15). The two
systems were identical otherwise. At the receiving end of the data line was a
solenoid that actuated a mechanical system for printing characters. In the later
systems, the solenoid was replaced with amplifiers and sensors. These were used
for process monitoring at remote distances. However, even though the electro-
/

mechanical switching devices were eventually replaced with much faster vacuum
tube circuits, and later still, with solid state devices, the current loop systems
were beginning to be replaced with the RS-232C systems.
396 Chapter 12 Computers and Automatic Control Systems

XMTR RCVR

1
1
1
Mechanical
20 mA, or 60 mA Actuator

Solenoid
Off
-u w-
On
rh

(a) Older Method

XMTR RCVR

Figure 12-15 Methods of using 20 mA and 60 mA current loops.

12-7.2 The RS-232C Standard

This standard was developed by the Electronic Industries Association (or EIA)
and was originally designed for use in communications involving printers, termi¬
nals (i.e., CRTs), and modems. The standard was originally referred to as the
EIA RS-232 standard but was almost immediately revised, consequently becom¬
ing known by the present designation of the RS-232C standard. This standard is
another serial data system. It’s interesting to note that even though serial trans¬
mission of data requires only two lines, one for the data and another for ground
return, the RS-232C standard specifies 25 pin connections in all for the connectors
at each end of the cable that must be used. The pins are identified in Table 12-2.
Figure 12-16 shows the pin configuration on the D-type connector commonly
used for RS-232 interfacing. In reality, most of the pins are not used for most
applications. Pins 1, 2, 3, and 7 are the only pins really needed in many installa¬
tions. As few as two pins, pins 3 and 7, can be used in some installations for
receiving data only. Figure 12-17 depicts a typical RS-232C installation between a
computer and a process station. The maximum baud rate for an RS-232C system
Section 12-7 Methods of Computer Interfacing 397

TABLE 12-2 EIA RS-232C INTERFACE


STANDARD SHOWING PIN CONNECTOR
ID

Pin Description

1 Protective ground
2 Transmit data
3 Receive data
4 Request to send
5 Clear to send
6 Data set ready
7 Signal ground (common return)
8 Received line signal detector
9 (Reserved)
10 (Reserved)
11 Not used
12 Secondary rec’d line signal detector
13 Secondary clear to send
14 Secondary transmitted data
15 Transmit signal timing
16 Secondary received data
17 Receiver signal timing
18 Not used
19 Secondary request to send
20 Data terminal ready
21 Signal quality detector
22 Ring indicator
23 Data signal rate detector
24 Transmit signal element timing
25 Not used

Figure 12-16 The 25-pin D-type con¬


nector.
o
GO
C/3
<uO
os_
Q.
<U
o
E

TJc
3
<U
3
C_
E
oo
c
<D<u
£
-O
aj
C/3
c
CSCJ
'E
3
E
E
o
o
a
o5
-a
U
(N
ro
r-i
i
oo
Cti

4/
k. —
3
0£ • —«
3
£ CJ

398
Section 12-7 Methods of Computer Interfacing 399

is 20,000 baud. That is, only 20,000 bits per second of data can be handled with
the serial interface.

12-7.3 The Centronics System

The Centronics system is really not an interface scheme used for control systems
as such. It is presented here only because of its extensive use in computer sys¬
tems and to make this discussion on interface systems complete. The Centronics
system is used primarily for interfacing printers to computers (see Figure 12-18).
It is a parallel interface that is probably the most popular system used today for
printer communications. It can transfer data at tremendously high rates com¬
pared to the serial interfaces just discussed, which accounts for the high-speed
printing rates that are common today. In Figure 12-18 we see two connector pin¬
outs illustrated. One is for the connector that goes into the printer device, while
the other is the connector that goes into the computer or CPU. The CPU’s con¬
nector is constructed in such a manner that it is not possible to mistakenly plug the
wrong end of the interconnecting printer cable into the computer’s printer connec¬
tor. The connector on the printer is a 36-pin connector containing eight lines that
carry the eight-bit wide word in parallel. The transmission of the binary data is
controlled by a computer-generated strobe signal. The flow control of the binary
data is done through handshaking; this is the turning on (causing a line to go high
with 1 bit) or turning off (causing a line to go low with a 0 bit) of the ACKNOWL¬
EDGE or BUSY leads, or a combination of both these lines. The connector on
the cable’s other end is a 25-pin connector similar to the D-type connector used
for the RS-232C system. The pins’ wiring connections, however, are radically
different. Also, the 25-pin connector for the RS-232C cable is a female connector,
whereas the 25-pin Centronics cable connector is a male connector.

12-7.4 The IEEE-488 General Purpose Instrumentation Bus

The International Electronics and Electrical Engineering Society standardized a


parallel interface scheme called the IEEE-488 General Purpose Instrumentation
Bus, otherwise known as the IEEE-488 GPIB. The system was actually devel¬
oped by Hewlett Packard Corporation for their instrumentation needs but was
later adopted and standardized by the IEEE organization. The 488 bus allows
two-way communications between as many as one dozen or so different instru¬
ments equipped to be attached to the system. However, the spacing of these
instruments from one another is rather limited, for reasons that were mentioned
earlier in our discussion on communication techniques. The circuitry needed for
this system is quite complicated requiring specially designed ICs for its implemen¬
tation. Basically, the 488-bus system allows instruments to transmit, receive, and
control data flow between themselves in a very ordered and error-free fashion. In
general, the 488 interface is comprised of three kinds of devices: (a) The talker:
This is a device designed to place data on the bus, allowing it to be transmitted to
05 cn
CD a) CD CD
c _c

■S -i
^ > _C •- C
"5 13.2
05 "O . 2 cd CD CD
o +->
CD c : CD $ ~ ID
o Q. O >•
CD k_
J= o
GO 4-»
>
4-J
“ x
CD cd
CD -a ~
CD
Q_
CD
a
fS
CD CD ~o 5
03 • CD o JZ > CD C '
o> 2 CD _ > CD CD
6 CD aX - o ~ «
C in o 0-0^
c c 4_ ' ,y a co ~0 CD 5
O -H CD o CD C ■M cl £ a.
CL
Q- o 05 03 CO O — *“* CD >
CD
CD _j co
uS CJ — E CO — c CJ "O "O
■- (D CD
o t; 05 22
CN c
o
5 10
Q. 03
Q)
JD _
co

CD
<D

^ c
CD
T3
C
c .E
T5 =D ^
c
UJ 1- -C c 4—1 > CD
C c
f“ r-
CD „ ^ cj
05 CD
CQ to <£ .§> - CD
si
o
o —
°
4-» c-
-C

aj O g s
CD
C ^
5 22
— c
Q-

"O
l_
O
o
05
co
<D
Q
k!
C/5 a)
•M °
co 03
<— LLI 1 r1 —I
:
CD
_q ci -a
>- CD c t «
-» S
£ "O -C
c £ °
u -o
CO
CD
C •” Z. U
o £ S Z c §
cccccccc
III
D
o
3
O 5 8 2
CD > O)
q| o co c 0
CJ
“o .2 05
c ”
CD X

co ?
J <1> ■D

QQ
<—csioo^Lncoi^co o CO
4—1
c CD
co
03
o <<<<<<<< >- D
C
05 QC hl-Hhhl-l-l- CO CN Q_
>

<<<<<<<< 3 m
~ o
w.
C
cd
CO h-
CO 03000000 CJ
<
00 *— S— o
E
ocT 8
»— D
E
o X>
O

o
T3
CD CJ
C
o CD c/3
3
2
Or-CNn^fintDrs oo
E r>*
05
CD c (NCNCNICSICNCNCNICM cn CN E 3 OC
o “O
(T d. CO
O c
CD
o

CD c
CM CD

CO >“ CO
ft i 2
C CD C D •<
JZ N—
■4-J *
c ^ CNICO^LncDI^OOOi CD
4—*
•a
c s
.E5 c o
. I-S
(V
o r»-*
• *■».
<u
c
c rs
o i =Q
r»»
o # •<
-cc
C/5
3
’5.
c
■<
?»«*
o

=
iD
u SJ
D
sz
H •v
U
X
u
1 D
<N
C
D ’lZ
u a.
3
OJC D
-C
£ ■*—<

400
Section 12-7 Methods of Computer Interfacing 401

one or more connected devices. As a rule, only one talker is allowed to talk at a
time, (b) The listener: This device is designed to accept data from the bus and is
allowed to do so through a system called handshaking. There can be more than
one operating listener on the bus at any given time, (c) The controller: This is a
device whose job is to issue commands to both the listeners and the talkers.
There may be more than one controller on the bus, but only one controller can
operate at a time.
The connector style, pin-out, and additional pin descriptions for the connec¬
tors used in this system are all shown in Figure 12-19. Because some of these

Pin
Data Lines
Number
D101 through D108 used for data
12 Shield
Control Lines 11 ATH
IFC = Interface Clear 10 SRQ
SRQ = Service Request 9 IPC
REN = Remote Enable 8 NDAC
7 NRFD
EOI = End or Indentify
6 DAV
Handshaking Lines 5 EOI
4 D104
DAV = Data Valid 3 D103
NRFD = Not Ready for Data 2 D102
NDAC = No Data Accepted 1 D101

Figure 12-19 The IEEE-488 General Purpose Interface Bus. Connector and illus¬
tration and pin-out information reproduced by permission of Black Box Corpora¬
tion, 1987 ed. of Black Box® catalog, Pittsburgh, PA.

devices must both receive and transmit data and yet not interrupt the flow of data
on the bus, a provision must be made in the connector design to allow for the
“stacking” of connectors, much like the stacking of appliance plugs at an AC
socket outlet. Further information on this system can be obtained from the refer¬
ences listed at the end of this chapter.
402 Chapter 12 Computers and Automatic Control Systems

12-8 PLUG-IN CIRCUIT BOARD INTERFACING METHODS

Up to this point, we have discussed methods for interfacing circuits with other
circuits that are external to the computer. When designing an automatic control
system to operate around a custom-built microprocessor-based system, the prob¬
lem develops of interfacing internal circuit boards with other circuit boards de¬
signed to operate directly with a mother board containing the CPU. This is a
custom systems design problem that is solved by using one of the several existing
circuit board interfacing standards. This is a discussion for advanced digital cir¬
cuit design, and we don’t get into it here at this level. However, it’s important to
be able to recognize the names of these interface standards if only for reference
purposes. The IEEE-696/S100 standard (sometimes referred to simply as the
S100 bus) is used extensively for designing plug-in circuit boards, as is the STD-
BUS system. The term bus merely refers to the circuit paths that exist between
two or more devices in a computer system that are used for the transferring of
data, control signals, and operating power. The S100 bus card is shown in Figure
12-20(a); the STD bus card is shown in Figure 12-20(b). Other systems exist also,
and their combined totals probably outnumber the two just mentioned. But you
can be assured that these other more numerous systems will simply be minor
variations of the SI00 and STD bus systems. As to which of the two systems is
the more popular, the SI00 bus would probably be the choice. While the STD bus
is a better defined system (it has a 56-pin bus with 8 data lines, 16 address lines, 22
control lines, and 10 power line connections), the S100 bus has 16 data lines, 24
address lines, 11 interrupts, and is capable of multiprocessing (i.e., it can support
the operation of two or more CPUs within the same system). Address lines carry

Note: Total of 100 Pin Connections


Note: Total of 56 Pin Connections
(50 on Each Board Side)
(28 on Each Board Side)

(a) (b)
All Dimensions in Inches

Figure 12-20 Two popular edge connector codes.


Review Questions 403

the data placement information so that the data get stored in the proper location
within a register or other computer-related storage area. Interrupts are signals
used to temporarily halt the operating system. These signals come from either
within the computer itself for system supervisory reasons, or they originate from
outside the computer, possibly coming from an I/O device indicating a completion
of data transfer.
Because of the increasing usage of computer systems and components in
automatic control systems, it makes sense for the design technician or engineer to
become familiar with these various bus systems. It’s becoming increasingly diffi¬
cult for mechanical engineers and technicians to remain purely mechanical, and
for electrical engineers and technicians to remain purely electrical. The area of
automatic controls almost demands that there be a cross-pollination of ideas and
information between these two fields in order to keep up with the increasing
complexity of control system development and design.

SUMMARY

Unquestionably, the single biggest change in automatic control design has been
the adaptation of computer control. This change has forced many of the control
designers to rethink their design philosophy and to learn new hardware and soft¬
ware technology. In some cases, especially in Chapter 12, you may not have seen
the relevancy of all the emphasis on computer technology with automatic con¬
trols. That’s easy to do, since there is so much to learn in this subject. People in
industry have had an especially difficult time with this. More often than not, they
have had to adapt and learn computer science on their own time because of the
magnitude of information they needed to consume. Then, having done that, they
have had to figure out how to integrate all that information into their particular
application. The use of programmable controllers in process control is just one of
these adaptations industrial users are having to cope with. Certainly, the com¬
munications industry has had to do the same thing. A good example of this is the
use of computer remote control in communications satellites orbiting the earth
and the remote control of extraterrestrial robots. Every one of these systems
requires some form of automatic control for its proper operation. But, in order to
understand the control’s system of operation, it’s necessary to understand how
the computer operates.

REVIEW QUESTIONS

12-1. Explain the various functions of the microprocessor chip. What does the
ALU do in this chip?
12-2. Explain the purpose of a data bus in a microcomputer.
404 Chapter 12 Computers and Automatic Control Systems

12-3. Explain the difference between RAM and ROM memory chips. Why is it
necessary to have both?
12-4. Explain why we can’t depend solely on floppy disks and hard disks for
memory in a computer. In other words, why can't we do away with RAM
chips entirely?
12-5. What is the purpose of the BIOS chips in a microcomputer?
12-6. Explain the difference between volatile and nonvolatile memory.
12-7. What is meant by the language level of a computer language? Why is
assembly language considered a lower level language as compared to
BASIC?
12-8. What is the function of a compiler? What is the basic difference between a
compiler and an interpreter?
12-9. Explain the function of a PLC. What is a PLC’s I/O?
12-10. Explain what is meant by Centronics connector, parallel connector, and
RS-232 connector.

REFERENCES

Artwick, Bruce A., Microcomputer Interfacing, Englewood Cliffs, N.J.: Pren¬


tice-Hall, Inc., 1980.
Boyce, Jefferson C., Digital Computer Fundamentals, Englewood Cliffs, N.J.:
Prentice-Hall, Inc., 1977.
Brey, Barry B., Microprocessor/Hardware Interfacing and Applications,
Columbus, Ohio: Charles E. Merrill Co., 1984.
Gaonkar, Ramesh S., Microprocessor Architecture, Programming, and Appli¬
cations, Columbus, Ohio: Charles R. Merrill Co., 1984.
Gilmore, Charles M., Introduction to Microprocessors, New York, N.Y.:
McGraw-Hill Book Company, 1981.
Gunn, Thomas G., Computer Applications in Manufacturing, New York, N.Y.:
Industrial Press, Inc., 1981.
Streitmatter, Gene A. and Vito Fiore, Microprocessors Theory & Applica¬
tions, Reston Va.: Reston Publishing Company, Inc., 1979.
Tocci, Ronald J. and Lester P. Laskowski, Microprocessors and Microcom¬
puters: Hardware and Software, Englewood Cliffs, N.J.: Prentice-Hall, Inc.,
1979.
Wagner, T. J. and G. J. Lipovski, Fundamentals of Microcomputer Program¬
ming, New York, N.Y.: Macmillan Publishing Company, 1984.
Answers
to Odd-Numbered
Problems
The following are the answers to the odd-numbered problems in the listed chap¬
ters. Chapter 1 is a review chapter; therefore there are no answers listed. Chap¬
ter 9 contains design problems; therefore, because of the variety and complexity
of solutions that are possible, no answers are offered there either.

CHAPTER 2:

2-1: 25.7 2-3: Desired


Set Point
Temp.

Error Signal
Output Temp.
-► Chiller
(Inside Chamber)
M

Temp. Sensor Feeding


Back Signal to Temp.
Controller
Temp. Controller
Acting as a Diff.
Amp. Producing
Error Signal

2-5: 20.68 in./gal. 2-7: a. 60 dB b. Not definable since one can’t calculate the
log of a negative number. Also, negative voltage gains are not definable, c.
23.2 d. -3.74 2-9: -3.02 dB

CHAPTER 3:
2.5
3-1: 107A 3-3: 6/s3 + 9/s2 - 1 Is 3-5: Assume that 10 = 1/r, then
5(0.1 + 1)
406 Answers to Odd-Numbered Problems

1
3-7: Assume 1/4 = a, then 3-9: Assume r = 3, then
^ + 0.25 5(35 + 1)
3-11: ?'2b) . 3-13: 3-15: f(t) = 22te~' 3-17: /(*) = 6.2e-3'
52 + to2 (5 + 0.5)2
3-19: lOsinlO/

CHAPTER 4:
j?
^out 1 + 543.33 x IQ"6
4-1: T.F. = 0.305 VDC/in. 4-3: r - 0.018 sec 4-5:
'in 1 + 5173.3 x 10-6
48 50(1 + 0.06»
4-7: T.F. = 4-9: T.F. =
5(1 + 0.0325) joj( 1 + 0.1jcu)( 1 + 0.008yoo)

CHAPTER 5:
EOUt T7 10 G\G2 + G-\
_ if AAA C '1. ^OUt _ 'out
5-1: = 35,000 5-3: 5-5:
E\in 1 'in 5+15 Em 1 + H(G,G2 + Gj)
5*out G\G2
5-7: 5-9: ^ = 3363
Siin //, + g3h2 Fin
1 + G\G2

CHAPTER 6:

1 1 I 1 IMI 1 1 1—rriTT] 1 1—1—r hit -1-1 1 1 1 rTT


6-1: Phase Angle Plot for All Values of K.
0
- -90

-180
TF = 125
40
Phase Angle (Deg)
TF = 17

< 20 TF = 10
QQ
-a
CD TF = 1
o 0

-20

-40

_1_1_i_1 1 1 l 1 _1_1 1 1 1 1 1 1 _1_1_1 1 1 1 1 1 _1_1_1 1 1 l 1 1


0.1 10 100
co (Rad/Sec)
Answers to Odd-Numbered Problems 407

Phase Angle (Deg)


Phase Angle (Deg)

0.1 1 10 100
co (Rad/Sec)
408 Answers to Odd-Numbered Problems

Phase Angle (Deg)


0.1 1 10 100
co (Rad/Sec)

-90

°
-18

Phase Angle (Deg)

0.1 1 10 100
co (Rad/Sec)
Answers to Odd-Numbered Problems 409

CHAPTER 7:

co (Rad/Sec)

7-3: The open-loop bandwidth is 13.5 r/s. The closed-loop bandwidth is 16 r/s. In
general, going to a closed-loop system usually increases a system’s bandwidth.
o
-90

-180

QJ
Q
jO)
CT)
C
<
CD
CO
cu

o.i 10 100
co (Rad/Sec)
410 Answers to Odd-Numbered Problems

7-7: The Nichols plot should resemble very closely the closed-loop curve of
Problem 7-1. 7-9: Approximately -20 dBA in gain reduction is needed for
stabilizing this system.

CHAPTER 8:

8-1: a. zeros: = -5 b. zeros: = -0.167


poles: = 0, -3.33, +0.33 poles: - —0.33, +1, —1
c. zeros: = none d. zeros: = -0.33
poles: = -10, -2 poles: = 0, -2, -0.1
e. zeros: = -4 /. zeros: = 5
poles: = -1.01, -0.04 poles: = 2, -0.167
8-3:
Amplitude

Time (Sec)

8-5: con = 7.746 r/s 8-7: ajn = 25 r/s


ojd = 7.484 r/s z = 0.125
% overshoot = 43.2% ±5% settling time = 0.96 5
number of osc. = 3.79
ojm — 24.6 r/s
8-9: Eq. (8-41) states the following: asymptote location pt. = 25

CHAPTER 10:

10-1: An analog automatic control system has the advantage of not having as
much of a resolution problem in set-point control as compared to a digital
automatic control system since there are no incremental (digital) settings to
Answers to Odd-Numbered Problems 411

contend with. Theoretically, an analog system is capable of an infinite degree of


resolution because of the infinite number of set-point opportunities. 10-3:
Because of the uncertainty existing with the output of a binary-type encoded
wheel or slide during the transition from one binary position to the next, the Gray
code was necessary in order to eliminate this uncertainty or ambiguity of
switching. The switching problem arises from the encoder’s sensors detecting
this change-over at slightly different times as the boundaries separating the
binary-generating areas sweep beneath the sensors, thereby creating a series of
random, unwanted binary number readings. 10-5: A bit map is a picture of an
image that is to be digitized showing the locations of the binary bits representing
the various light intensities on that image. 10-7: Adaptive control is an
automatic control system that, through its feedback, can make corrections to an
operating system. The one distinguishing feature of this type of control system as
compared to any other automatic control system, is that an adaptive system
responds to the environmental changes surrounding the main automatic control
system. The purpose of an adaptive control system is to prevent an automatic
control system from destroying itself in its attempt to make an error correction
due to some sudden or unanticipated change within its operating environment.
10-9: An optical encoder can be made to have a much larger travel displacement
to accommodate larger measurand values as compared to an LVDT. Because of
this, much greater resolution is possible. Any attempts to increase or to amplify
the displacement of an existing LVDT using mechanical means usually results in
back-lash and/or hysteresis.

CHAPTER 11:

11-1: The three major categories of data transmission are: (1) analog, (2) digital,
and (3) carrier transmission.
Analog transmissions are characterized by a continuously changing or
modulated voltage or current whose amplitude is a linear representation of the
measurand being sensed at the transmission site.
Digital transmissions are characterized by a transmission of rectangular or
pulse-like voltages or currents whose amplitudes, positions, widths, or frequen¬
cies of occurrence represent the measurand at the transmission site.
Carrier transmissions are characterized by the systematic interruptions of a
transmitter carrier wave, the duration and frequency of the interruptions
containing the transmitted data or information.
11-3: FM transmissions are more often used for data transmissions as compared
to AM transmissions because of less electrical interference present with FM
transmissions. This is because most electrical interference, either man-made or
natural, is of an AM nature rather than of an FM nature. 11-5: In an FM system
the carrier’s frequency is modulated by the transmitted data whereas in a PM
system, the electrical phase existing between the carrier and the transmitted data
412 Answers to Odd-Numbered Problems

is varied. There is little advantage in using the one method over the other,
although, it is believed that PM may have a slight advantage over FM in electrical
noise immunity.
11-7: The four basic forms of pulse modification are: (1) pulse amplitude
modulation, (2) pulse width modulation, (3) pulse position modulation, and (4)
pulse code modulation.
Pulse amplitude modulation is characterized by its varying pulse heights
containing the transmitted data. Pulse width modulation is characterized by its
varying pulse widths containing the transmitted data. Pulse position modulation
is characterized by the positions or locations of each pulse relative to a reference
location, the position containing the transmitted data. Pulse code modulation is
characterized by the presence or absence of each pulse sent in groups
representing coded binary numbers.
11-9: An advantage of serial data transmission is that it requires fewer
transmission channels (a minimum of two, one for the signal and one for ground
return) as compared to a parallel transmission system which may require a
minimum of 4 (8 being more typical), depending on the “word” length being
transmitted. Another advantage is that serial transmissions can be transmitted
over great distances as compared to parallel transmissions. A disadvantage of a
serial transmission is that it is very slow compared to a parallel transmission.
Parallel transmissions have the great advantage of being very fast as
compared to serial transmissions. However, their disadvantages are, very limited
transmission distances, greater complexity in circuitry and many more
transmission channels required.

CHAPTER 12:

12-1: The function of the microprocessor chip is to control the flow of traffic, so to
speak, of all the data that needs to be processed. The chip contains data registers,
instruction sets, binary arithmetic functions for data conversions, data address
registers, and program counters to properly manipulate the data’s flow and
processing.
The function of the ALU, or arithmetic-logic unit, is to perform the
necessary math functions and logic needed for processing the input data.
12-3: RAM (random-access memory) refers to a type of memory chip that acts as
a temporary storage location of data. However, RAM is volatile; that is, the
memory capability is lost when power is removed from these chips. The ROM
(read-only memory) chip, on the other hand, is a non-volatile chip that contains
data or information that can be read from it. That is, the information can be
extracted at any time; however, it is not possible to store user data within this
chip. ROM chips are used for storing programs that can be used at any time when
desired. When power is removed, the ROM’s contents still remain intact. 12-5:
BIOS (basic input-output system) chips contain a computer’s operating
Answers to Odd-Numbered Problems 413

commands that control the computer’s many input/output ports which, in turn,
allow the computer to send and receive data from peripheral devices attached to
these ports. 12-7: The “level” of a computer language refers to that language’s
ability to perform a command or series of commands with a minimum number of
command statements and symbols. The higher level languages require fewer
statements and symbols as compared to the lower level languages that require
many statements. However, the higher level languages tend to lose versatility in
their applications as compared to the lower level languages. 12-9: The function
of a PLC (programmable logic controller) is to act as a sequencing or logic device,
and sometimes as a data processor, for an industrial machine used in a
manufacturing facility. The I/O on a PLC refers to the input/output ports which
connect to, and receive data from, the machine that the PLC is controlling.
Typical devices attached to a PLC’s I/O are, the machine’s control switches,
sensors attached to the machine, and annunciators.
APPENDIX A
Rectangular! Polar
Polar!Rectangular
Conversions and How
to Handle Math Routines
Using These Forms
When working with electronics and in automatic control system design, it’s
necessary to be able to convert quickly from rectangular expressions to polar
expressions and back again. The reason why we have to contend with both sys¬
tems of expression is because adding and subtracting requires one expression
form while multiplication and division requires another form. Also, depending on
the application, one form is probably easier to understand than the other. For the
purposes of automatic control applications, you should know both systems and be
able to switch from one to the other.
Figure A-l illustrates the trigonometric relationships between the two sys¬
tems for locating a point, m. Point m can be given the position (A,jB) using the

4
m

jB

Figure A-l Graphical relationship


between rectangular and polar coordi¬
nates.

414
Appendix A Rectangular!Polar Polar/Rectangular Conversions 415

rectangular coordinate system, or it can be given the position (R L 6) using the


polar coordinate form of system. To convert from the rectangular to the polar
form, use the following conversion equations:

Rectangular —» Polar: R = VA2 + B2 (A-l)


±B
6 = arctan —— (A-2)

EXAMPLE -
Convert (2 + j5) to polar form.

Solution:
Use Eq. (A-l) to find R:

R = V22 + 52
= V29
= 5.385

Use Eq. (A-2) to find 6:

5
6 = arctan -
0 = arctan 2.5
= 68.2°

The final answer is expressed as 5.385 Z. 68.2°.

To convert from the polar form to the rectangular form, use the following
equations:

Polar Rectangular: A = RcosO (A-3)


B = Rsind (A-4)

EXAMPLE -
Convert 2.87 A -56.8 to the rectangular equivalent.

Solution:
Use Eq.(A-3) to find A:

A = (2.87)(0.548)
= 1.573

Use Eq. (A-4) to find B:

B = (2.87)(—0.837)
- -2.402

The final answer is expressed as (1.573 — j'2.402).


416 Appendix A RectangularlPolar Polar!Rectangular Conversions

Handling mathematical manipulations using the two coordinate forms can be


easy if you choose the right form for the right math process. For instance, you
will want to use the rectangular expressions only for addition and subtraction.
You will want to use the polar forms only for multiplication and division. You can
violate these rules, but the math can really become a problem. The following are
examples of how to perform these four math functions:

EXAMPLE -
Add (4.7 L 30°) to (-6.4 - 73.7).

Solution:
Since we are adding, we will want to use all rectangular forms. Therefore, convert
the one polar form that was given to rectangular form, then add the two expressions
as you would normally do to any algebraic expression. Converting (4.7 Z. 30°) to
rectangular form:

A = 4.7cos30° (Eq. A-3)


= 7.070
B = 4.7sin30° (Eq. A-4)
= 2.350

Therefore, (4.7 Z 30°) - (7.07 + 72.35)


Now adding the two rectangular values:
7.07 + 72.35
(+) -6.34 - 73.70
Our answer-> 0.73 — 71-35

EXAMPLE --
Subtract (3.87 + 78.33) from (1.19 -76.98).

Solution:
No conversions are necessary, since both given expression are already in the desired
rectangular form.

Solution:
1.19- 76.98
(-) 3.87 + 78.33
Our answer-» - 2.68 -715.31

EXAMPLE -
Multiply the preceding two expressions together.
Appendix A Rectangular!Polar Polar!Rectangular Conversions 417

Solution:
First, we must convert both expressions to polar coordinates. (We won’t show the
conversion math this time, just the results.):

1.19 - 76.98 —> 7.081 Z -80.3°


3.87 + 78.33 -> 9.185 Z 65.1°

The rule for multiplying polar expressions is this: The R values of the expressions are
multiplied as usual, but the angle values (i.e., the 6s) are added. Therefore,

(7.081 Z -80.3°) x (9.185 Z 65.1°) = 7.081 x 9.185


- 65.039 - 80.3° + 65.1° - -15.2°

Our answer —» 65.039 Z -15.2°

EXAMPLE -
Divide 64.8 Z —124.0°
4.99 Z 23.6°

Solution:
The rule for division of polar expressions is this: The R values of the expressions are
divided as usual; however, the angle values (i.e., the 0s) are subtracted. Therefore,

64.8 - 4.99 = 12.986


(-124.0°) - (23.6°) = -100.4°

Our answer^ 12.986 Z —100.4°


APPENDIX B
A Computer Program
for Plotting Transient
Waveforms

Being able to obtain a visual image of the transients generated by a closed-


loop feedback system is a great step toward understanding the system itself.
Provided that you have the equation describing the transient response curve, the
BASIC program that follows will give you such a plot. This program was de¬
signed to run on an AT IBM computer. However, it will most likely run on any
IBM personal computer with little or no modification. If it is run on a compatible
IBM, the BIOS chip(s) have no BASIC program and therefore no IBM graphics
capability which this program requires. Therefore, you must use either GW
BASIC, BASIC, or BASIC A along with a graphics routine enabling you to print
graphics to your hi-res monitor. The program will not function without the graph¬
ics routine.
Follow these instructions:

1. Type the code in Figure B-l, being very careful to duplicate every character
and symbol used.

5 SCREEN 0
6 REM This program is called PLOTFUN.BAS. It was originally published in
7 REM MACHINE DESIGN magazine, Feb. 26, 1987, J. Perkins, author. The
8 REM program has been modified to suit this particular application.
10 PRINT "PROGRAM PLOTFUN . . . PLOTS FUNCTION ON SCREEN"
20 PRINT " "
30 NL = 1
40 OPEN "FUNCTION.BAS" FOR OUTPUT AS #1
50 PRINT "ENTER THE EQUATION TO BE PLOTTED, BEGINNING WITH *Y = * AND HAVING"
51 PRINT "EQUATION EXPRESSED IN BASIC NOTATION AND A FUNCTION OF ’ T’.“
52 PRINT "WHEN COMPLETED, HIT <ENTER>."
55 INPUT A$
60 L = 990 + 10*NL

Figure B-l BASIC code for PLOTFUN.BAS. (Reprinted with permission from MA¬
CHINE DESIGN magazine, Vol. 59, number 4.)

418
Appendix B A Computer Program for Plotting Transient Waveforms 419

65 NL = NL + 1
70 LNUM$=STR$(L)
90 A$ = LNUM$ + " " + A$
100 PRINT #1,A$
120 PRINT #1, "2000 RETURN-
130 CLOSE #1
140 CHAIN MERGE "FUNCTION.BAS",200,ALL,DELETE 1000-2000
200 REM
210 INPUT "ENTER TIMESCALE;TMAX",TMAX
220 INPUT "ENTER VERTSCALE;YMIN,YMAX",Y1,Y2
230 Xl=0:X2=TMAX:DELT=TMAX/640
240 REM
250 REM PLOT THE SECTION ON THE SCREEN
260 CLS:SCREEN 2: KEY OFF
270 XS=639/(X2-X1):YS=197/(Y2-Y1)
280 REM
290 PSET(0,0)
300 LINE-(0,199)
310 LINE-(639,199)
320 LINE-(639,0)
330 LINE-(0,0)
340 REM
345 REF=200*Y2/(Y2-Y1)
350 PSET(0,REF)
360 LINE-(640,REF)
370 PSET(0,REF-(Y*YS))
380 REM
390 T=0
400 XX=Y
410 GOSUB 1000
420 UU=DELT*XS + UU
430 VV=REF-(Y*YS)
440 LINE-(UU,VV)
450 T=T+DELT
460 IF T<(TMAX+DELT) GOTO 400
470 END
480 REM
1000 REM
2000 REM

Figure B-l Cont.

2. After you are certain the code has been copied correctly, SAVE the pro¬
gram. Then, type RUN. You will be prompted to type the equation which
must be entered in the form, Y = f(T). For example:

Y = 1 + 1.05 * 2.718^-30 * T) * SIN(105 * T - 100)

3. After typing your equation, hit (ENTER).


4. You will then be asked to enter the maximum time (TMAX) scale you think
you will need to completely plot your equation along the horizontal time axis
so that it has time to settle out.
5. Next, you will be asked to make two entries on the same line separated by a
comma. You must enter the minimum vertical scale needed to begin the
curve (YMIN). Typically, this would be 0. You will also be asked to enter
the maximum scale (YMAX) needed to contain the curve’s highest point. If,
420 Appendix B A Computer Program for Plotting Transient Waveforms

for example, you think 4 is enough for your YMAX, then you must make
your complete YMIN, YMAX entry as: 0,4. For the preceding example
curve, a TMAX of 0.25 and a YMIN,YMAX of 0,2 works well.
6. After hitting (RTN) (or (ENTER), whichever the case may be), the curve
will become plotted in high-resolution mode. The horizontal line at the top
of the graph represents your maximum vertical scale value, YMAX, while
the lowest horizontal line represents your minimum vertical scale value,
YMIN. The far right-hand vertical line on your graph is located at your
TMAX value.
7. Following the plot, you must enter SCREEN 0 to go back to low res if you
wish to run another plot. Also, your soft key line will be OFF at this time.
Therefore, also enter KEY ON to return that line, if desired.
8. To run another plot, you must restore line 1000 and line 2000 in your pro¬
gram back to its original state. Do this by typing 1000 REM (RTN) and 2000
REM (RTN) to restore the program. Check your entries by LISTing the
program.

If you had entered the example equation given in step 2, you would obtain a
plot similar to the one shown in Figure B-2.
This is obviously a “bare-bones” program and not too “friendly.” You may
want to modify it to have it perform in the manner you wish.
OK

Figure B-2 Typical results from using PLOTFUN.


APPENDIX C

Computer Programs
for Finding Cube Roots

The two BASIC programs listed following have a rather interesting applica¬
tion. The program listed in Figure C-l uses an iterative process in determining the
roots for cubic equations. This is a process developed by Newton and is surpris¬
ingly accurate and efficient when done on a computer. The BASIC program is
called FINDROOT.BAS. It can be used to find all three roots if desired. It does
take a few moments to use, and a certain amount of patience is required to apply
it. The program will not find roots less than -1000. The equation must be entered
in one of the program lines along with the equation’s first derivative. The process
of finding the first derivative of a polynomial is rather easy. Here is how it is
done. The exponent associated with the variable is multiplied by the variable's
coefficient. The exponent’s value is then reduced by one.
Here are some examples:

Find the first derivative of 5y6.


Solution: First derivative — 6 • 5y6-1 = 30y5.

Find the first derivative of 123v.


Solution: First derivative = 1 • 123v1-1 = 123.

Find the first derivative of \4p~4.


, Solution: First derivative = -4 • 14p~4~] = -56p~5.

Find the first derivative of —56 (which can be written as — 56jc°).


Solution: First derivative = 0 • -56x0_1 = 0.
This tells us that the derivative of any constant will be 0.

421
422 Appendix C Computer Programs for Finding Cube Roots

Find the first derivative of the function y = 3x1 2 3 4 5 6 + lx — 2.


Solution: First derivative = yi = 6x + 7.

Now that we know how to find the first derivative of a polynomial expres¬
sion, what does it do for us? The first derivative is nothing more than the equation
that allows us to calculate the slope of a line passing through the points at a point
of tangency to the original curve. Placing those points into the derivative, yl, (or
y' as it is sometimes called) and solving, gives us the slope of that line of tan¬
gency. FINDROOT.BAS lets you blindly guess at the first root of an equation. It
then calculates the slope of the line based on your first guess and also determines
the point of intersection of that slope line with the x-axis. The computer then
plugs the value of that intersection point back into the original equation and
recalculates f(x) to see if the resultant value of/(x) becomes smaller. If it does,
the process is repeated until two successive values of /(x) are separated by less
than 1 x 10-6. When the separation becomes that small, the value of x is declared
a root for /(x).
Obviously, there are three roots for a cubic equation. As a result, you may
want to guess at other values of x to see if the computer will converge on the other
remaining ones. By the way, if the program doesn’t converge within 50 attempts,
this, most likely, is an indication that either the program is trying to converge on
an imaginary root, or your guess was so far off that it is going to take more than 50
tries to bring in an answer. In either case, you will receive an error message
stating “— slope is zero”, or “— no convergence in XXX tries”. If you should
determine one root and don’t wish to spend the time guessing for the others, you
can then go to the next program, ROOTS.BAS, (listed following) to solve for the
others. That program is more direct in finding the remaining roots.
FINDROOT.BAS is listed in Figure C-l. The following are the instructions
for using the program:

1. Enter the program into your computer.


2. While still in BASIC or BASICA, type line number 8400 and enter the
equation that you wish to find the roots of. Be sure to begin entry with
“F = ”.
3. Next, type line number 8420 and enter the first derivative of the preceding
equation, beginning the entry with “FI = ”.
4. Type RUN. You will be asked to guess at the first root. Pick any number no
less than —1,000. If you do, you will receive an error message.
5. All iterations will be displayed showing the converging process on the one
root. Once that root is finally displayed, you can either guess at another or
go to the next program for calculating the remaining roots in the case of
cubic solutions.
6. In order to get out of the program, you must depress the CTRL-BREAK
keys simultaneously.
Appendix C Computer Programs for Finding Cube Roots 423

10 REM This program is called FINDROOT.BAS and is based on using the Newton-
11 REM Raphson method. This program was originally obtained from BASIC
12 REM PROGRAMS FOR SCIENTISTS AND ENGINEERS, by A. R. Miller, Sybex, 1981.
13 REM D6 DX delta x
14 REM El% ERMES% error flag
15 REM F FX function
16 REM FI DFX derivative of function
17 REM F0% FALSE% zero
18 REM H2 SMALL small number
19 REM M5% MAXL% maximum loops
20 REM T0% TRUE% not false
21 REM T1 TOL tolerance
22 REM end of identifiers
23 REM
30 T1 = .000001
40 H2 = IE -15
50 M5% = 50
60 F0% = 0
70 T0% = NOT F0%
80 INPUT "First guess ";X
90 IF (X < -1000) THEN 9999
100 GOSUB 8000
110 PRINT
120 IF (El% = F0%) THEN PRINT "The solution is ";X
130 GOTO 80
8000 REM start of Newton’s method
8010 El% = F0%
8020 FOR L% = 1 TO M5%
8030 XI = X
8040 GOSUB 8400
8050 IF (ABS(Fl) > H2) THEN 8090
8060 PRINT "ERROR - slope is zero"
8070 El% = T0%
8080 GOTO 8160
8090 D6 = F/Fl
8100 X = XI - D6
8110 PRINT "X = ";XI;", FX = "; F;" DFX = ";F1
8120 IF (ABS(D6) < = ABS(T1 * X)) THEN 8160
8130 NEXT L%
8140 PRINT "ERROR - no convergence in "; M5%;" tries"
8150 El% = T0%
8160 RETURN : REM from Newton’s method
8400 REM Enter cubic equation here, beginning with "F = ".
8420 REM Enter derivative of above equation here, beginning with "FI =
8430 RETURN
9000 END
9999 PRINT "ERROR -- you have exceeded the negative input range"
10000 PRINT "for your guess. Please RUN program again and guess again."
10010 END

Figure C-l BASIC code for FINDROOT.BAS. (Reprinted with modifications and permis¬
sion from BASIC PROGRAMS FOR SCIENTISTS AND ENGINEERS, A. R. Miller, Sybex
Inc., copyright 1981, Alameda, Calif.)

As an example, let’s assume that you have to find the roots of the following
cubic equation:

F = 3x3 + lx2 — 4x — 3
424 Appendix C Computer Programs for Finding Cube Roots

We will need the first derivative of this equation in order to run the program, so
let’s determine that also.
Tj = 9x2 + 14* - 4
Placing these two expressions into lines 8400 and 8420, respectively, and remem¬
bering to use BASIC notation while we are doing it, let’s use “500” as our first
guess for the first root determination. Figure C-2 shows the results of this guess.
First guess ? 500
X — 500 , FX = 3.76748E+08 DFX = 2256996
X — 333.0755 , FX = 1.116287E+08 DFX = 1003112
X — 221.7931 , FX = 3.307493E+07 DFX = 445830.8
X 147.606 , FX = 9799821 DFX = 198150.1
X — 98.14941 , FX = 2903548 DFX = 88069.86
X — 65.18073 , FX = 860242.1 DFX = 39145.27
X — 43.2051 , FX = 254841.2 DFX = 17400.99
X — 28.55988 , FX = 75478.49 DFX = 7736.841
X — 18.80416 , FX = 22344.21 DFX = 3441.626
X — 12.31182 , FX = 6607.528 DFX = 1532.593
X — 8.000481 , FX = 1949.329 DFX = 684.076
X — 5.150901 , FX = 572.1066 DFX = 306.8986
X — 3.286746 , FX = 165.9891 DFX = 139.2387
X 2.094627 , FX = 46.90404 DFX = 64.81195
X — 1.370933 , FX = 12.4023 DFX = 32.10818
X — .9846669 , FX = 2.712423 DFX = 18.51146
X — .8381403 , FX = .3311205 DFX = 14.05628
X — .8145835 , FX = 8.031607E-03 DFX = 13.37609
X — .813983 , FX = 4.529953E -06 DFX = 13.35888

The solution is .8139827


First guess ?
1LIST 2RUN 3LOAD" 4SAVE" 5CONT 6,"LPT1 7TRON 8TROFF9KEY 0SCN

Figure C-2 Results from running FINDROOT.

We see that the first root is 0.8139827. It’s interesting to study the displayed
iterations used in finding this root. This will help you in understanding how the
program works. To find the remaining roots, we could continue our guessing or
we could go on to the next program. Let’s continue guessing just to see what our
results will be. This time, let’s start with a very negative number such as
“—500”. Using this as our guess, we come up with a second root value of
-2.69074. If we continued guessing, we would find our third root to be
-0.45676. This was found using a guess of “-2”.
As in the case of the previous program in Appendix B, FINDROOT.BAS
lacks conveniency for entering data. You can revise the code to make it more
suitable for you if you like. This program can also be used for determining the
roots of many other kinds of equations, as long as you also enter the proper
derivatives for those equations.
Another program for finding roots is called ROOTS.BAS. It allows you to
find the remaining roots of any cubic equation regardless of whether those roots
are real or imaginary. The program assumes that you know, within a value of
Appendix C Computer Programs for Finding Cube Roots 425

±0.5, the value of the first root. If your first root is outside that range value, the
program will tell you in an error message. The listing for ROOTS.BAS is given in
Figure C-3. The program is run by doing the following, assuming that you have it
entered properly into your computer:

1. Type RUN.
2. You will first be prompted to enter what you believe is the first root. (This
first root value could come from FINDROOT.BAS.) Hit (ENTER).

1 REM This program is called ROOTS.BAS


2 REM This program calculates the other two roots of a cubic equation when
3 REM given the first root, XI
5 INPUT "XI ROOT = ";X1
10 INPUT "X-CUBED COEFFICIENT = ";S3
20 INPUT "X-SQUARED COEFFICIENT = " ; S2
30 INPUT "X COEFFICIENT = " ; SI
40 INPUT "EQUATION’S K-VALUE = ";K
50 PRINT S3
60 PRINT S2
70 PRINT SI
80 PRINT K
90 REM begin synthetic division
100 SYN1 = XI * S3
110 SYN2 = S2 + SYN1
120 SYN3 = XI * SYN2
130 SYN4 = SI + SYN3
140 SYN5 = XI * SYN4
150 SYN6 = K + SYN5
155 IF (ABS(SYN6) > .05) THEN 500
160 REM S3 now a-term for quadratic formula
170 REM SYN2 now b-term for quadratic formula
180 REM SYN4 now c-term for quadratic formula
190 BSQ = SYN2~2
200 AC4 = SYN4 * S3 * 4
210 HLD = BSQ - AC4
220 IF HLD >= 0 THEN 240
230 IF HLD < 0 THEN 300
240 RT = HLD''. 5
250 X2 = -SYN2/(2*S3) + RT/(2*S3)
260 X3 = -SYN2/(2*S3) - RT/(2*S3)
270 PRINT "X2 = ";X2
280 PRINT "X3 = ";X3
290 END
300 HLD = AC4 - BSQ
310 RT = HLD''. 5
340 PRINT "X2 = “;-SYN2/(2*S3)
350 PRINT "+J";RT/(2*S3)
360 PRINT "X3 = ";-SYN2/(2*S3)
370 PRINT "-J";RT/(2*S3)
380 END
500 PRINT "ERROR - first root, XI, not accurate enough. Retry program"
510 fRINT "by entering RUN again and incrementing ’first guess’ to XI"
520 PRINT "by a value of plus or minus .05."
540 END

Figure C-3 BASIC code for ROOTS.BAS.


426 Appendix C Computer Programs for Finding Cube Roots

3. Next, you will be asked to enter the coefficients associated with each vari¬
able in the cubic equation you are trying to solve. If a variable is missing,
simply enter “0”. Each entry is followed by (ENTER).
4. Your display will then show a listing of your coefficients that you have
entered followed by the calculated remaining two root values.
5. To rerun the program, simply enter RUN.

To test the program, let’s follow through with the example that we had begun
in the first program, FINDROOT.BAS. Let’s enter the first solved root of
0.8139827, along with the coefficient information of our cubic equation. The dis¬
play should look like Figure C-4. If you wish, you can also enter the second and
third roots found from the first program. You should find that these results com¬
pare with the displayed results in Figure C-3.
RUN
XI ROOT = ? .8139827
X-CUBED COEFFICIENT = ? 3
X-SQUARED COEFFICIENT = ? 7
X COEFFICIENT = ? -4
EQUATION’S K-VALUE = ? -3
3
7
-4
-3
X2 = -.456576
X3 = -2.69074
Ok

1LIST 2RUN 3LOAD” 4SAVE 5CONT 6,"LPT1 7TRON 8TROFF9KEY OSCN

Figure C-4 Results from running ROOTS using results from FINDROOT.

The ROOTS.BAS program will also return complex or imaginary values of


roots. For example, find the roots for the cubic equation, y = 0.125x3 - 0.25x2 +
0.5x + 3 given that one of its roots was already found to be -2. ROOTS.BAS will
return the display seen in Figure C-5. We see that the other two roots are
2 + 72.828427 and 2 - J2.828427.
Appendix C Computer Programs for Finding Cube Roots 427

RUN
XI ROOT = ? -2
X-CUBED COEFFICIENT = ? .125
X-SQUARED COEFFICIENT = ? -.25
X COEFFICIENT = ? .5
EQUATION’S K-VALUE = ? 3
. 125
-.25
.5
3
X2 = 2
+J 2.828427
X3 = 2
-J 2.828427
Ok

1LIST 2RUN 3LOAD" 4SAVE" 5CONT 6,"LPT1 7TRON 8TROFF9KEY OSCN

Figure C-5 Results of a complex root solution using ROOTS.


APPENDIX D

Euler’s Equation

The purpose of this appendix is to present a proof for Eq. (8-22) in Section
8-6.1, which is called Euler’s equation. This equation states that ejd =
cos0 + jsin#. This is a very handy identity to know in coping with many of the
mathematical expressions that have a mixture of trigonometry expressions and ex
expressions. This mixture occurs quite often in automatic control theory presen¬
tations.
To begin the proof, we first look at one of the math tools we need for this
job. We rely on what is called the Maclaurin’s Series. This is an expression that
allows you convert a trigonometric expression into an algebraic expression. The
Maclaurin’s Series states:

/« = m + rmx + m*i
2!
+ rm*i
3
+ !
mmx"
n\

As you can see, in order to use the preceding expression, you must be able to take
multiple derivatives of terms. If you are unsure as to how this is done, you should
obtain a good elementary text on differential calculus and review these proce¬
dures. You will find these procedures to be very straightforward, especially for
sine and cosine expressions. For example, let’s convert sin x into an algebraic
expression.

f(x) = sin x when x = 0 /(0) = 0


fix) = cos x (first deriv.) /'(0) = 1

428
Appendix D Euler’s Equation 429

f'\x) = -sin x (second deriv.) /"(0) - 0


/"'(*) = —cos x (third deriv.) /'"(0) = -1
/4(x) = sin x (fourth deriv.) f\ 0) = 0
/5(x) = cos x (fifth deriv.) f\ 0) = 1
Substituting the values from our table into the Maclaurin’s Series, we get:

_ ... (0)x2 ( l)x3 (0)*' (O*5


sin x 0 + (l)x + Hrr- + + +
2! 3! 4! 5!
x5
120

Using the same approach for cos x, we would get:

X2 X4 X6
cosx=l-- + -- - +

And, if we did the same for ex we would get:

. x^ x
l+X + yj- + — + . . .

Now, let x = jx in the preceding ex expression, remembering that

j = j
j2 = -1
P = ~j
P = 1
j5 = j
Then,

jjP x4
eJx = l + jx
3!

Rearranging the terms, we get:

x2 x4 X3 X5

1 “ 2! + 4! + j x “ 3! + 5!

and cos x 1 —> sin x

Therefore, eJx = cosx + j sin x.


Similarly, we can prove that e~Jx = cosx — j sin x.
AUTOMATIC
CONTROL
SYSTEMS
AND COMPONENTS
Index

A Auto-pilot for boat, control system


for, 275-80
Absolutely unstable system, 236 Average power, 48 (see also Power)
AC:
circuits, 22-28
servo, 108-20 B
servomotor, torque curve (figure),
122 Backlash, 288
Altitude controller, aneroid barome¬ Bandwidth, 57, 193
ter, 69-71 Bang-bang servo, 299-304
AM demodulator, 131 designing control systems for,
Ampere, defined, 2 299-304
Amplification (see Gain) Block diagram, 66-67
Anemometer, hot-wire, 100 explained, 141-45
Angle of departure, root locus mathematics, 141-56
method, 247 terminology, 144-45
Apparent power, 48 (see also Power) Blocks:
Armstrong oscillator, 49 (see also cascading, 145, 147-48
Oscillators) combined reductions, 151-56
Asymptote location point, 248 interchanging, 151
Audio frequency response, audio Bode diagram:
amplifier, 56 determining stability, 178-83

433
434 Index

Bode diagram: (cont.) Closed-loop:


elements, 171-74, 184 peak frequency, how calculated,
general construction rules, 162-63, 223
169 peak gain, how calculated, 223
graphing complex functions, 175— plotting advantages, 195
78 versus open-loop system (figure),
straight-line method, accuracy of 187
(tables), 175 Color code, resistor (table), 8
Bode, H. W., 160 Compensating circuit, 133
Breakaway point, 250 Complex:
Break point (see Corner frequency) number, 45
s-plane, 231-33
Components, automatic control
system, 66-67
c control of, 67
error determining, 67
Capacitive: feedback control, 67
microphone, 97 input of, 66
reactance, 40-42 output of, 67
Capacitor, 28-31 Computer plotting, transient re¬
dielectric constants (table), 29 sponse curve, 418-20
Capacitors: Conditionally stable system, 293
parallel, 38 Control:
series, 31 receiver, 124, 125, 126, 127
series-parallel, 31 synchros, 123
Carbon microphone, 95-96 transformer, synchro, 126-27
Centroid (see asymptote location transmitter, 124, 125, 126, 127
point) Control Systems, designing, 258-304
Characteristic equation, 233-37 auto-pilot for boat, 275-80
first-order, 241 bang-bang servo, 299-304
poles of, 235 hydraulic press, 295-99
roots of, 235 oscillograph, 289-95
second-order, 243 robotic cart steering, 258-75
third-order, 247 valve positioner, 280-89
zeroes of, 235 Controller:
Chart, Nichols, 195-203 (see also derivative, 254
Nichols chart) digital, 295-99
Circuit: integral, 253-54
parallel, 13, 30, 34 proportional, 252-53
series, 10, 31, 33 Controller types, 252-55 (see also
series-parallel, 15, 31, 34 Controller)
vibrator reed, 129 Corner frequency, 168
Circular mil, 6-7 Coupling, magnetic, 35
Index 435

CR, 124, 125, 126, 127 Domain, s-plane, defined, 80


Critically damped, 214 (see also Drift, DC amplifier, 109
Damped) Dynamic microphone, 97
Crossover times, transient response
curve, 216-19
Current:
E
instantaneous, 24, 46-48
RMS, 46-48
Effective value, 46-47
Current direction, how determined, 4
Electric charge, 2
Current flow, explanation, 1
Electrical time constant, 117
Current law, Kirchhoff s, 18-19
Electromagnetic:
CX, 124, 125, 126, 127
sensing, 101
Cycle, 24
spectrum, 101-2
Encoder:
optical linear, 297
D photocell, 90
rotary magnetic, 93-94
Damped:
Euler, Leonard, 231
frequency, 214-19
Euler’s equation, 231
sinusoidal wave, 207
Damped, critically, 214
Damping:
factor, 212 F
inertial, 118
ratio (see Damping factor) Factor:
viscous, 114, 115, 265-67 damping, 212
DC: power, 49
amplifier drift, 109 Farad, 29
servo (see Servomotor) Feedback, 61-66
Dead band, 261-62, 288 eliminating loops, 147, 148
Decibel, 55 negative, 64, 72, 145
Decibel, sound, 95 positive, 61, 72, 145
Demodulator, 130-32 Feedback control, 61
AM, 131 FET transistor, 52
full-wave, 131 Flowrate sensing, 99
half-wave (figure), 131 FM discriminator, 131
Derivative controller, 254 Free-wheeling speed, 114
Dielectric constant (see Relative Frequency, 24
permittivity) damped, 214-19
Digital controller, 295-99 defined, 24
Diode, 52 maximum gain, 189
Discriminator, FM, 131 natural resonant, 212
Divider rule, voltage, 19-20 resonant, 49-51
436 Index

Frequency response, 54-56 Inductive reactance, 42-43


audio amplifier, 56 Inductor, 31-34
negative feedback in, 65 Inertial damping, 118
human ear, 55 Inertially damped servomotor, 118—
F (s), defined, 80 20
Full-wave demodulator, 131 Instantaneous:
current, 24, 46-48
power, defined, 24
G voltage, 24, 46-48
Integral controller, 253-54
Gain, 55, 56-57, 145 Integrated circuit, 53
maximum system, 189 Inverter (see Modulator)
unity, 179
Gear:
ratio, 121 J
train, 120-22
Jitteriness, 181-82, 184
J-operator, 45-46
H

Half-wave demodulator (figure), 131


Henry, 32, 34 K
Hot-wire anemometer, 100
Hunting (see also, Damping factor), Kirchhoff’s laws, 17-19
303, 304 for current, 18-19
Hydraulic press, designing control for voltage, 17-18, 21
system for, 295-99
Hysteresis, 301, 303
L
I Lagging phase shift, 59
Laplace, Pierre Simon, 74
IC, 53 Laplace transform:
IC, FM radio, 103 cosine function, 79
IDSM, 118-20 first-order system, 79
Imaginary number, 45 parabola, 79
Impedance, 32, 44 ramp function, 79
Inductance: second-order system, 79
parallel, 34 sine function, 79
series, 33 step input, 79
series-parallel, 34 Laplace transforms:
Induction motor: compared to logarithms, 75
torque curve (figure), 112 equivalent math operations using
two-phase, 109-10 (table), 81
Index 437

how used, 74-83 MOSFET transistor, 52


pairs (table), 79 Motor, stepper, 134-37
versus calculus, 75-78
LC circuit, 50
Leading phase shift, 59 N
Light sensing, 106-8
Linear variable differential trans¬ Natural resonant frequency, 212
former, 90 Negative feedback, 64, 72, 145
Linearity, 57 Network:
Locus, 239 phase-lag, 133-34
Logic gates, 53 phase-lead, 132-33
Lumen, 106 Nichols chart:
LVDT, 90 how used, 195-203
defined, 90 paper, 195
transfer function, 90 NPN transistor, 52
Number:
complex, 45
M imaginary, 45

Magnetic coupling, 35
Manometer, 100 o
Margin, phase, 181
Maximum: Ohm, 5, 6
gain frequency, 189 Open-loop system, 61
system gain, 189 Optical linear encoder, 297
Mechanical time constant, 117 Oscillator, Armstrong, 49
Microfarad, 28 Oscillators, 49-51
Microhenry, 34 Oscillograph, designing control sys¬
Microphone: tem for, 289-95
capacitive, 97 Overshoot, 114
carbon, 95-96 Overshoot, in transient response
crystal (see Microphone, piezo¬ curve, 221-22
electric)
dynamic, 97
electromagnetic (see Microphone, P
dynamic)
piezoelectric, 96-97 Parallel:
transfer function for (see Transfer capacitance, 30
function, sonic) inductance, 34
Millihenry, 34 resistance, 13-14
Mil, circular, 6-7 Peak value, 3-dB, 189-90
Minicomputers, 375 Period, 24
Modulator, 128-30 Permeability, 32
Moment arm, 112 Permittivity, 29
438 Index

Permittivity, relative, 29 R
Phase, 57-61
angle, 26-28 Radian, 25-26
-delay devices, 134, 295 RAM, 375-76
-lag network, 133-34 Rate generators, 122-23
-lead network, 132-33 Ratio:
margin, 181 gear, 121
shift, voltage (see Gain)
lagging, 59-60 Reactance, 39-44
leading, 59 capacitive, 40-42
Phasors (see J-operator) inductive, 42-43
Photocell encoder, 90 Read-only-memory, 376
Photoconductive sensor (see Photo¬ Real power, 48
resistive sensor) Regenerative receiver (see Arm¬
Photoemissive sensor, 108 strong oscillator)
Photoresistive sensor, 106 Relative permittivity, 29
Photovoltaic sensor, 108 Relay operated system, 302-3 (see
Picofarad, 28 also Servo, bang-bang)
Piezoelectric: Resistance:
materials (table), 97 how calculated, 9-20
microphone, 96-97 parallel, 13-14
Plane, right-hand, 241 series, 10-12
PNP transistor, 52 series-parallel, 15-17
Polar form, 46 Resistivity, 6-7
Polarity, how determined, 4-5 common materials (table), 6
Position sensing, 87-90 defined, 6
Positive feedback, 61, 72, 145 Resistor, 8-20
Potentiometer, 87-90 color code (table), 8
Potentiometer, configurations (fig¬ tolerance, 9
ure), 88 wattage, 9
Power, 48-49 Resolvers, 127, 128
apparent, 48 Resonant frequency, 49-51
average, 48 Response frequency, 54-56
factor, 49 audio amplifier, 56
instantaneous, defined, 24 human ear, 55
real, 48 Response time, transient response
supply regulation, 20-22 curve, 222-23
Pressure sensing, 94-95 Response, transient, 205
Proportional controller, 252-53 Right-hand:
plane, 241
rule, 23
Q RMS:
current, 46-48
Quadratic formula, 242 voltage, 46
Index 439

Robotic cart steering, designing con¬ Series-parallel:


trol for, 258-75 capacitance, 31
ROM, 376 inductance, 34
Root Locus method, 230-52 resistance, 15-17
angle of departure, 247 Servomechanism, 108
first-order functions, 239-41 Servomotor, 108-20
second-order functions, 241-46 Servo, bang-bang, 299-304
third-order functions, 246-52 Servo-generators (see Rate gener¬
Root Locus, plotting, 239-52, 268, ators)
279, 287 Settling:
Root mean square (see RMS) oscillations, transient response
Rotary magnetic encoder, 93-94 curve, 220-21
Routh-Hurwitz stability test, 237-38 time, transient response curve,
R-C transient behavior, 37-38 219-20
R-L transient behavior, 38-39 Sluggishness, 181-82, 184
Sound sensing, 95-99
Spectrum, electromagnetic, 101
Speed ratio (see Gear ratio)
s Speed, free-wheeling, 114
5-plane, complex, 231-33
Seebeck Effect, 104 5-plane domain, defined, 80
Self-inductance (see Inductance) Stability test, Routh-Hurwitz, 237
Sensing: 38
element, 85-86 Stable system, conditionally,
flowrate, 99 293
Sensing: Stall torque, 114
electromagnetic, 101 Steady state condition, 205
light, 106-8 Step input:
position, 87-90 defined, 207
pressure, 94-95 transient response, 207
sound, 95-99 unit, 208
temperature, 104-6 Stepper motor, 134-37
velocity, 91-94 Stepper motor, bit pattern (table),
Sensor: 135
defined, 85-86 Summing point, 67-68, 144
photoconductive (see Sensor, pho¬ Summing point, moving, 149-50
toresistive) Switching transistors, 129
photoemissive, 108 Sychro control transformer, 126—
photoresistive, 106 27
photovoltaic, 108 Synchromechanisms, 123-27
Series: Synchros:
capacitance, 31 control, 123
inductance, 33 torque, 123
resistance, 10-12 System, open-loop, 61
440 Index

T rack and pinion gear, 122


radio receiver, 103, 137
Tachometer (see Rate generators) rate generator, 123, 138
Take-off point, 144 resistive velocity sensing, 91
Temperature sensing, 104-6 servomotor, 111-20
Thermocouple, 104, 106 sonic, 98, 137
3-dB: standard Bode form, 165
down points, 57 stepper motor, 137, 138
peak value, 189-90 synchromechanism, 125
Time constant: thermistor, 104
electrical, 117 thermo transducer, 137
mechanical, 117 thermocouple, 106, 138
Time-dependent expressions (see VDSM, 114-17, 138
also Time domain), 75 velocity transducer, 137
Time domain (see also Time-depen¬ Transformer, 34-37
dent expressions), 80, 83 air-core, 35
Torque, 111 iron-core, 35
constant, 115-16, 278 turns ratio, 35-36
receiver, 124, 125 variable core, 35
synchros, 123 Transformer power ratings, 36-37
transmitter, 124, 125, 126, 127 Transient behavior, 37-39
TR, 124, 125 R-C, 37-38
Transducer, defined, 85-86 R-L, 38-39
Transfer function, 68, 72 Transient response, 205
CR or TR synchro, 138 Transient response curve, 215-31
CX or TX synchro, 138 computer plotting, 418-20
demodulator, 132 cross-over times, 216-19
determining stability, 236 over-shoot, 221-22
flowrate, 101, 137 response time, 222-23
gear train, 121, 138 settling time, 219-20
IDSM, 118, 120, 138 settling time oscillations, 220-21
LVDT, 90 Transient waveform components,
modulator, 130 230
phase-lag network, 134, 138 Transient response to step input, 207
phase-lead network, 133, 138 Transistor:
photocell FET, 52
transducer, 138 MOSFET, 52
velocity transducer, 137 NPN, 52
photoemissive sensor, 108 PNP, 52
photoresistor, 107 Transistors, switching, 129
photovoltaic sensor, 108 Turns ratio, transformer, 35-36
potentiometer, 88-89, 137 Two-phase induction motor, 109-10
pressure sensing, 95, 137 TX, 124, 125, 126, 127
Index 441

u drop, 10
law, Kirchhoffs, 17-18, 21
Unit step input, 208 source, ideal versus actual, 3-4
Unity gain, 179 Voltage:
Unstable system, absolutely, 236 instantaneous, 24, 46-48
RMS, 46
Volt, defined, 2-3
V

VA, 49
Valve positioner, designing control
w
system for, 280-89
Wattage, resistor, 9
Variable core transformer, 35
Waveform:
VDSM, 114-17
amplitude, 24
Vectors, 45
cycle, 24
Velocity:
defined, 24
constant, 116
frequency, 24
sensing, 91-94
period, 24
Vibrator reed circuit, 129
Viscous:
damped servomotor, 114-17
damping, 114, 115, 265-67 z
Voltage:
divider rule, 19-20 Zener diode, 52

You might also like