100% found this document useful (1 vote)
715 views

Suits BH Electronics For Physicists An Introduction

Electronics

Uploaded by

Strahinja Donic
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
715 views

Suits BH Electronics For Physicists An Introduction

Electronics

Uploaded by

Strahinja Donic
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 340

Undergraduate Lecture Notes in Physics

Bryan H. Suits

Electronics
for Physicists
An Introduction
Second Edition
Undergraduate Lecture Notes in Physics

Series Editors
Neil Ashby, University of Colorado, Boulder, CO, USA
William Brantley, Department of Physics, Furman University, Greenville, SC, USA
Matthew Deady, Physics Program, Bard College, Annandale-on-Hudson,
NY, USA
Michael Fowler, Department of Physics, University of Virginia, Charlottesville,
VA, USA
Morten Hjorth-Jensen, Department of Physics, University of Oslo, Oslo, Norway
Michael Inglis, Department of Physical Sciences, SUNY Suffolk County
Community College, Selden, NY, USA
Barry Luokkala , Department of Physics, Carnegie Mellon University, Pittsburgh,
PA, USA
Undergraduate Lecture Notes in Physics (ULNP) publishes authoritative texts
covering topics throughout pure and applied physics. Each title in the series is
suitable as a basis for undergraduate instruction, typically containing practice
problems, worked examples, chapter summaries, and suggestions for further reading.
ULNP titles must provide at least one of the following:
• An exceptionally clear and concise treatment of a standard undergraduate subject.
• A solid undergraduate-level introduction to a graduate, advanced, or non-standard
subject.
• A novel perspective or an unusual approach to teaching a subject.
ULNP especially encourages new, original, and idiosyncratic approaches to physics
teaching at the undergraduate level.
The purpose of ULNP is to provide intriguing, absorbing books that will continue to
be the reader’s preferred reference throughout their academic career.
Bryan H. Suits

Electronics for Physicists


An Introduction
Second Edition
Bryan H. Suits
Department of Physics
Michigan Technological University
Houghton, MI, USA

ISSN 2192-4791 ISSN 2192-4805 (electronic)


Undergraduate Lecture Notes in Physics
ISBN 978-3-031-36363-4 ISBN 978-3-031-36364-1 (eBook)
https://doi.org/10.1007/978-3-031-36364-1

1st edition: © Springer Nature Switzerland AG 2020


2nd edition: © The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer
Nature Switzerland AG 2023

This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse
of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

Physicists rely heavily on electrical instrumentation to measure physical phenomena.


There was a time, not long ago in history, when it was normal for a physicist to
routinely design and construct their own electrical instrumentation. While there are
exceptions, in general, this is no longer the case; design and construction of most
electronic instrumentation is left to specialists and engineers. However, a physicist,
whether an experimentalist or theorist, will need some understanding of the basics of
electronics and how the equipment works in order to fully understand and evaluate
the results of measurements, and possibly in order to troubleshoot and make simple
repairs. Much of this understanding must first come from a knowledge of the language
of electronics—the lexicon as well as the symbolic representation of circuits using
schematic diagrams. This book was written for use as an undergraduate physics text
with such future use in mind. The details necessary for quality engineering design
are generally excluded here in favor of a basic and practical understanding of what
is going on.
A second utility of electronics in the physics curriculum is that many ideas and
problem-solving strategies show up that are also used in other areas of science.
Indeed, electrical analogies are often used as an explanation for situations that have
little to do with electronics. One prime example of cross-use is the appearance of
imaginary numbers. Imaginary numbers show up in all areas of science and engi-
neering that include periodic signals. In addition to electronics, those areas include
studies of vibration, including seismometry, acoustics, optics, radio and radar, and
even “brain waves.” It is important to understand how to use and interpret imagi-
nary numbers since, after all, by definition no real measurement will ever give you
an imaginary result. Other overlapping topics include resonance, solutions of linear
equations, and the use of linearized models.
Each electronics instructor will have their own idea about what is most important.
That cannot be helped. What is presented here represents my priorities. Based on
my perspective and experience as a physicist and teacher of electronics, I chose
introductory topics and problems that I find most interesting and potentially useful.
There is, however, more in this book than can possibly be covered in a single semester

v
vi Preface

course. Anyone using the book as a course textbook, in whole or in part, should feel
free to skip those topics that do not match their interests.
Any electronics course would be expected to include a laboratory component. In
fact, some of the material here originated as background material for such a laboratory
experience. It is recognized that each instructor has their own laboratory priorities,
and possibly a limited or specialized supply of laboratory equipment for such a
purpose. It is hoped that the presentation here is written in a manner suitable for use
in either the laboratory or classroom. The first obvious examples of laboratories that
appear here are the sections related to the Wheatstone and Kelvin bridges, both of
which originated primarily as laboratory exercises. Along with the material in the
last chapter, some practical experience using an embedded microcontroller—that is,
some programming—is definitely useful. Programming varies from device to device
and is not really “electronics,” and so it is only included here in a very general way.
The book includes three broad categories of electronics. Chapters 1–5 cover
passive linear electronics, Chaps. 6–11 look at nonlinear and active devices including
diodes, transistors, and op-amps, and Chaps. 12–14 consider the basics of digital
electronics and simplified computers.
This text originated as weekly handouts and laboratory write-ups for a course
designed primarily for second-year university physics students. The level of the
material here is appropriate for students who have successfully learned the material
in calculus-based introductory electricity and magnetism as well as mathematics up
to the first course in integral calculus. As the extent of the material in the handouts
expanded over the years, it got to the point that some students started referring
to the handouts as “the book.” That reference served as one motivating factor to
formalize those course notes into a full volume. I thank those students for providing
that inspiration, and I hope this proves to be a useful exercise for both of us.
The revised edition clarifies some of the discussion, adds some additional end-
of-chapter problems, and includes a new chapter, Chap. 15, about quantum-based
approaches. While there are no simple electronic components currently available for
quantum logic, the topic follows from the digital logic already discussed. The simi-
larities and differences between classical transistor-based electronics and quantum
approaches are emphasized without getting into details of how quantum devices
might be implemented.

Houghton, MI, USA Bryan H. Suits


Contents

1 The Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Voltage and Current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Simple Devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Kirchhoff’s Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Resistors in Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.5 Resistors in Parallel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.6 Effective Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.6.1 Resistors in Parallel–Notation . . . . . . . . . . . . . . . . . . . . 10
1.7 Solving Circuits with Circuit Reduction . . . . . . . . . . . . . . . . . . . . 11
1.8 Solving Circuits with Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.8.1 Branch and Mesh Currents . . . . . . . . . . . . . . . . . . . . . . . 13
1.8.2 Example—Using Kirchhoff’s Laws . . . . . . . . . . . . . . . 14
1.8.3 Nodal Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.9 The Ideal Current Source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.10 The Ground and Common Connections . . . . . . . . . . . . . . . . . . . . . 20
1.11 Multiple Sources—The Superposition Theorem . . . . . . . . . . . . . 21
1.12 Electrical Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
1.13 Additional Application—The Kelvin-Varley Divider . . . . . . . . . 23
1.14 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2 Additional Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.1 Thevenin and Norton Equivalents . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.1.1 Determining the Thevenin and/or Norton
Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.1.2 How Is This Used for Circuit Reduction? . . . . . . . . . . 32
2.1.3 Equivalent for an Infinite Array of Resistors . . . . . . . . 33
2.2 The Wheatstone Bridge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.2.1 Wheatstone Bridge “Hieroglyphics” . . . . . . . . . . . . . . . 36
2.3 The Reciprocity Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.3.1 Example—R-2R Ladder with Sources . . . . . . . . . . . . . 37

vii
viii Contents

2.4 Delta-Y Conversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40


2.5 The Kelvin Bridge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.5.1 Additional Application—Resistivity
of Lamellae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
2.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3 Complex Impedances . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.1 What is a Linear Device? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.2 Some Vocabulary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.3 Passive Linear Circuit Elements with Two Leads . . . . . . . . . . . . . 53
3.4 Idealized Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.5 RC and L/R Time Constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.5.1 RC Time Constant Example . . . . . . . . . . . . . . . . . . . . . . 57
3.6 Capacitors and Inductors with Sinusoidal Sources . . . . . . . . . . . . 58
3.7 Superposition and Complex Impedances . . . . . . . . . . . . . . . . . . . . 60
3.8 Series and Parallel Capacitors and Inductors . . . . . . . . . . . . . . . . 63
3.9 Comments About Complex Arithmetic . . . . . . . . . . . . . . . . . . . . . 65
3.10 Solving Circuits Using Complex Impedances . . . . . . . . . . . . . . . . 66
3.11 AC Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3.12 Condenser Microphones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.13 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4 More on Capacitors and Inductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.1 Real Capacitors and Inductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.2 Measuring Capacitors and Inductors . . . . . . . . . . . . . . . . . . . . . . . 76
4.3 Capacitive Position Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
4.4 A Simple Circuit for Measuring Inductors . . . . . . . . . . . . . . . . . . 78
4.5 Switched Capacitor Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
4.6 Charging a Capacitor Efficiently . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.7 Mutual Inductance and Transformers . . . . . . . . . . . . . . . . . . . . . . . 82
4.8 The Dot Convention for Transformers . . . . . . . . . . . . . . . . . . . . . . 84
4.9 Inductive Position Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.10 RLC Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
4.11 Cable Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
4.11.1 Cable Impedance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
4.11.2 Signal Speed in a Cable . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.11.3 Impedance of Finite Cables . . . . . . . . . . . . . . . . . . . . . . 92
4.12 Capacitor and Inductor Labels . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.13 Duality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
4.14 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
Contents ix

5 The Laplace Transform . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99


5.1 The Transform . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
5.2 Laplace Transform Example 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
5.2.1 Method I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
5.2.2 Method II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
5.3 Laplace Transform Example 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5.4 Laplace Transform Example 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.5 Comment on Partial Fractions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
5.6 Poles and Zeros . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
5.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
6 Diodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
6.1 Semiconductor Diodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
6.2 Diode Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
6.2.1 Piece-Wise Linear Diode Models . . . . . . . . . . . . . . . . . 119
6.2.2 An Analytic Model for the Semiconductor Diode . . . . 120
6.3 Solving Circuits with Diodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
6.3.1 The Ideal Diode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
6.3.2 Graphical Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
6.4 Diode Ratings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
6.5 Diode Capacitance and Response Time . . . . . . . . . . . . . . . . . . . . . 130
6.6 Specialty and Other Diodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
6.7 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
7 FETs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
7.1 Junction Field Effect Transistors . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
7.2 Circuit Analysis with a JFET . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
7.2.1 Example 1—Determine Circuit Components . . . . . . . 138
7.2.2 Example 2—Determining the Operating Point . . . . . . 139
7.3 The FET A.C. Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142
7.3.1 Example—Determine Transistor Parameters . . . . . . . . 143
7.4 FET Amplifier Configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
7.4.1 Example—Gain for Common Drain Amplifier . . . . . . 146
7.4.2 Example—Impedances for Common Gate
Amplifier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
7.5 The Ohmic Region . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
7.6 MOSFETs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
7.7 Additional Application—Dynamic Memory . . . . . . . . . . . . . . . . . 150
7.8 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
8 Bipolar Junction Transistors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
8.1 BJT D.C. Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
8.2 BJT A.C. Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
x Contents

8.2.1 BJT Large Signal Example—Graphical


Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
8.2.2 Single Supply Operation . . . . . . . . . . . . . . . . . . . . . . . . . 162
8.2.3 Solutions from Parameters . . . . . . . . . . . . . . . . . . . . . . . 163
8.3 BJT Amplifiers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
8.3.1 Example—Common Emitter Amplifier . . . . . . . . . . . . 168
8.3.2 Example—Common Collector Amplifier . . . . . . . . . . . 170
8.3.3 Using the Saturation Region . . . . . . . . . . . . . . . . . . . . . . 172
8.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
9 More on Amplifiers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
9.1 Miller’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
9.2 Two-Transistor Configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
9.2.1 The Cascode Configuration . . . . . . . . . . . . . . . . . . . . . . 179
9.2.2 The Darlington Pair . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
9.2.3 Complementary Symmetry Amplifier
(“Push–Pull”) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
9.2.4 Differential Amplifier . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
9.2.5 Current Mirror . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
9.2.6 Silicon Controlled Rectifiers (SCR) and Triacs . . . . . . 184
9.3 Connecting Amplifiers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
9.3.1 Impedance Matching . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
9.4 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
10 The Ideal Op-Amp . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
10.1 Ideal Op-Amp Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
10.2 Linear Op-Amp Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
10.2.1 Example 1—Buffer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
10.2.2 Example 2—Inverting Amplifier . . . . . . . . . . . . . . . . . . 192
10.2.3 Example 3—Non-inverting Amplifier . . . . . . . . . . . . . . 193
10.2.4 Example 4—Difference Amplifier . . . . . . . . . . . . . . . . 193
10.2.5 Example 5—Summing Amplifier . . . . . . . . . . . . . . . . . 194
10.2.6 Example 6—Integrator . . . . . . . . . . . . . . . . . . . . . . . . . . 195
10.2.7 Example 7—Low-Pass Filter . . . . . . . . . . . . . . . . . . . . . 197
10.2.8 Example 8—Instrumentation Amplifier . . . . . . . . . . . . 199
10.2.9 Example 9—A Capacitive Sensor for Smaller
Values of Capacitance . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
10.2.10 Example 10—Negative Resistor . . . . . . . . . . . . . . . . . . 201
10.2.11 Example 11—Constant Current Source . . . . . . . . . . . . 202
10.3 Other Op-Amp Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
10.3.1 Example 12—Non-linear Element in Feedback . . . . . 204
10.3.2 Example 13—Ideal Diode . . . . . . . . . . . . . . . . . . . . . . . 205
10.3.3 Example 14—Peak Follower . . . . . . . . . . . . . . . . . . . . . 206
10.3.4 Example 15—Log Amplifier . . . . . . . . . . . . . . . . . . . . . 207
10.3.5 Example 16—Absolute Value Circuit . . . . . . . . . . . . . . 208
Contents xi

10.4 More Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209


10.5 Less than Ideal Difference Amplifiers . . . . . . . . . . . . . . . . . . . . . . 210
10.5.1 Finite Input Resistance and Gain . . . . . . . . . . . . . . . . . . 210
10.5.2 Finite Frequency Range . . . . . . . . . . . . . . . . . . . . . . . . . 212
10.5.3 Small Signals and Drift . . . . . . . . . . . . . . . . . . . . . . . . . . 213
10.6 Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
10.7 The Transconductance Amplifier . . . . . . . . . . . . . . . . . . . . . . . . . . 215
10.8 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
11 Non-linear Uses of Op-Amps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
11.1 Limited Output Range . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
11.2 The Op-Amp Comparator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
11.2.1 Example 1—Low-Level Warning . . . . . . . . . . . . . . . . . 224
11.2.2 Example 2—Pulse Generator . . . . . . . . . . . . . . . . . . . . . 225
11.2.3 Example 3—Simple Oscillator . . . . . . . . . . . . . . . . . . . 226
11.2.4 Example 4—A Voting Circuit . . . . . . . . . . . . . . . . . . . . 227
11.2.5 Example 5—Sine to Pulse Train Converter . . . . . . . . . 227
11.2.6 Example 6—Zero Crossing Detector . . . . . . . . . . . . . . 228
11.2.7 Example 7—Pulse Conditioner/Lengthener . . . . . . . . . 229
11.3 Using the Comparator for Feedback . . . . . . . . . . . . . . . . . . . . . . . . 230
11.3.1 Automatic Gain Control Amplifier . . . . . . . . . . . . . . . . 230
11.4 Putting Pieces Together . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
11.4.1 Simple Phase Sensitive Detector . . . . . . . . . . . . . . . . . . 232
11.5 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
12 Digital I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
12.1 Boolean Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
12.2 Useful Rules and Theorems for Boolean Algebra . . . . . . . . . . . . 237
12.3 Digital Logic Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
12.4 Combinations of Digital Logic Gates . . . . . . . . . . . . . . . . . . . . . . . 240
12.4.1 Example 1—Solving with Boolean Algebra . . . . . . . . 240
12.4.2 Example 2—Solving with a Truth Table . . . . . . . . . . . 241
12.4.3 Example 3—Solving Both Ways . . . . . . . . . . . . . . . . . . 242
12.5 Equivalent Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
12.6 Gates Versus Logic Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
12.7 Decoders and Encoders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
12.8 Multiplexing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
12.9 Flip-Flop Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
12.10 Edge-Triggered Flip-Flops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
12.11 A Directional Electric Eye . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
12.12 Combinations of Flip-Flops . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
12.12.1 Shift Register . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
12.12.2 Binary Counter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
xii Contents

12.13 Other Non-logical Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . 259


12.13.1 Very Short Pulse Generator . . . . . . . . . . . . . . . . . . . . . . 259
12.13.2 Oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
12.14 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
13 Digital II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
13.1 Binary and BCD Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
13.1.1 Binary Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
13.1.2 BCD Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
13.1.3 Hexadecimal and Octal Notation . . . . . . . . . . . . . . . . . . 267
13.2 Other Weighted Binary Codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
13.2.1 The 4221 Code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
13.2.2 2 of 5 Codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
13.3 Non-weighted Codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
13.3.1 Gray Code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
13.3.2 The ASCII Code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
13.4 Bar Codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
13.4.1 Interleaved 2 of 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
13.4.2 UPC Codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
13.5 Some Numeric Code Conversions . . . . . . . . . . . . . . . . . . . . . . . . . 274
13.5.1 Base-2 Binary to Gray Code . . . . . . . . . . . . . . . . . . . . . 274
13.5.2 Gray Code to Base-2 Binary . . . . . . . . . . . . . . . . . . . . . 275
13.5.3 Decimal to Gray Code . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
13.5.4 BCD to Binary Conversion . . . . . . . . . . . . . . . . . . . . . . . 276
13.5.5 Binary to BCD Conversion . . . . . . . . . . . . . . . . . . . . . . . 277
13.6 Digital to Analog Conversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 278
13.6.1 The 1-Bit D/A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 278
13.6.2 A Summing D/A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
13.7 Analog to Digital Conversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
13.7.1 Voltage to Frequency Conversion . . . . . . . . . . . . . . . . . 281
13.7.2 Timing Schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282
13.7.3 Search Schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282
13.7.4 Analog to Gray Code Conversion . . . . . . . . . . . . . . . . . 283
13.8 Quantization Noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
13.9 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
Reference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
14 Calculators and Computers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
14.1 Adding Base-2 Numbers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
14.2 Two’s Complement Arithmetic . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
14.3 A Simple Arithmetic Logic Unit (ALU) . . . . . . . . . . . . . . . . . . . . 290
14.4 Base-2 Multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
14.5 Some Recursive Computations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 295
14.5.1 Compute 1/x . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 296
14.5.2 Compute (1/x)½ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 296
Contents xiii

14.5.3 Compute x½ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297


14.5.4 Compute tan(x) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298
14.5.5 Compute K(k) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298
14.6 Communications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
14.7 Tri-state Outputs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
14.8 Simplified CPU . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300
14.9 Other Uses for Tri-state Devices . . . . . . . . . . . . . . . . . . . . . . . . . . . 300
14.9.1 Measuring a Small Capacitance . . . . . . . . . . . . . . . . . . . 300
14.9.2 Charlieplexing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303
14.10 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
15 Quantum Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 305
15.1 Reversible Gates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306
15.2 Schematics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
15.3 Tensor Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308
15.4 Quantum Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
15.5 The Qubit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
15.6 Entanglement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
15.7 The Nature and Future of Quantum Computations . . . . . . . . . . . 318
15.8 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320

Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
Chapter 1
The Basics

In this first chapter, the basic concepts, definitions and results from electricity and
magnetism that are applicable to electronics are reproduced, along with some addi-
tional definitions more specific to electronics. Simple circuits involving resistors,
wires, and power sources are represented by schematic diagrams and then solved
using Ohm’s law and Kirchhoff’s laws. Simple theorems are developed for series
and parallel resistors which can be used to simplify a circuit, leading to a simpler
solution method.
Historically, physicists have often been required to design and create their own
specialized electronic circuits as part of their vocation. During the last few decades
that has become less and less of a requirement. One may ask, then, why it is a standard
practice to have modern physics students study electronics at all. Certainly, there is
no expectation that most will be designing state of the art electronic circuitry. There
are electrical engineers who are usually better equipped to do that. The goal here,
however, is not to teach the level of detail necessary for engineering work, but to give
an appreciation for what is involved, some understanding of how circuits work, and
at the same time, focus on some basic skills and techniques that show up in other
areas of science. Such an understanding is an essential part of figuring out what goes
into and what comes out of both experimental and theoretical scientific studies.
Hence, while venturing through this material try to keep an eye out for the “big
picture” which goes beyond the specifics of the electronics. Of course, the first key
to understanding is to have some knowledge of what you are talking about. That
knowledge starts by learning basic definitions, the lexicon, and the basic rules that
apply. The second key is to be able to apply logic and mathematics to deduce an
end result. With those thoughts in mind, the starting point here becomes the simplest
definitions and rules, in this case for electronics. It is expected that at least some of
this should be review. Following that, new results will be gradually introduced that
can be used to solve and understand more advanced electronic circuits.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 1


B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_1
2 1 The Basics

1.1 Voltage and Current

In electronics “voltages” and “currents” are of primary concern.


Voltage—what is measured by a voltmeter. Voltage arises from a source of electromotive
force (emf). If the source involves forces that are conservative, then the voltage is directly
related to an electrostatic potential difference. In the case of some time-dependent emfs, such
as arise due to Faraday’s law of induction, the emf is not associated with an electrostatic
potential but its effect is similar.1 In electronics, a distinction between the two is rare. They
are both simply referred to as “voltages.”

Voltages are measured in “volts,” abbreviated V.2

Current—a net movement of electric charge past a particular point measured over time.
Current is measured with an Ammeter.3

Currents are measured in “amperes,” or “amps” for short, abbreviated A, where 1 ampere =
1 coulomb of charge per second.

An applied voltage can be thought of as being somewhat like a force (or a pressure)
and the current as being a motion—a flow—in response to that force.
In mechanics, one is often given a force and then attempts to find the motion
(i.e., the response) and/or is given a motion and attempts to deduce the force. In
electronics, the same is done with voltages and currents. To “solve” a circuit is to
find the various voltages and currents—that is, to predict what you would measure
with a voltmeter and/or an ammeter.
Volts and amps are Standard International (SI) units. An appropriate standardized
prefix can be added to indicate powers of ten. A list of some of the more common of
these can be found in the Appendix. For example, 1000 V = 1 kV, 10−3 A = 1 mA,
etc. On some older electronics devices one may see “mm” (literally milli-milli-)
instead of “µ” (micro-) to indicate 10−6 . In addition, some older electronic devices
may be labeled using “m” or “M” instead of “µ” to mean micro- and hence “mm”
(or “MM”) becomes pico- (10−12 ). These peculiarities will have to be resolved from
the context or by direct measurement. This rather confusing font issue resulted from
the fact that Greek fonts were not readily available to (U.S.) manufacturers until
the latter part of the twentieth century. Fortunately, this prefix confusion does not
occur very often for more recently manufactured electronic components.4 To avoid
confusion, the use of standard SI prefixes is encouraged.

1 In other contexts, EMF may be used to stand for electromagnetic fields, which, incidentally, may
be a source of an electromotive force.
2 Named after Alessandro Volta (1745–1827), an Italian scientist credited with many discoveries

leading to the electric battery.


3 Named after French mathematician and scientist, André Marie Ampère (1775–1836), who, among

many important discoveries in physics, did many measurements to quantify currents.


4 In some fields of study, it is still the practice to use the ambiguous prefix “mc” for “micro,” though

that use is not usually seen in electronics.


1.2 Simple Devices 3

Since electronics deals with the relationship between voltage and current their
ratio will show up quite often. The ratio of 1 V/1 A = 1 u, where the upper-case
Greek omega (u) stands for ohms, the SI unit of electrical resistance.5
As a practical matter, the “u” symbol is often omitted in electronics when its
presence is clearly implied by circumstance. That practice will often be used in this
text. Also, in some circumstances decimal points can may be hard to see. If that is a
concern, prefixes may be used to substitute for a decimal point. For example, a value
of 4.7 ku (4700 u) may appear as “4.7 k” and sometimes as “4k7.” When there is
no prefix, the letter R (or r) is sometimes used for a decimal point. Hence, a value
of “4R7” would be 4.7 u. Less often, units other than ohms may be missing when
their presence is clearly implied by circumstance.

1.2 Simple Devices

Any electronic device will have one or more connection points, called “leads”
(pronounced lēds, as in “the captain leads their troops into battle”). A device with
one lead will not be useful as part of an electronic circuit, hence only devices with
two or more need be considered.
The electrical characteristics of a device are specified by the relationship between
the currents into (or out of) each lead corresponding to a given voltage between the
leads—there is a “device rule” that specifies the relationship. Two generic devices are
illustrated in Fig. 1.1. By convention, currents for each lead are usually considered
positive going into the device. As will be seen in more detail later, “net current in”
= “net current out” so, for the two-lead device shown, I 1 = –I 2 ≡ I and so only
the current I going through the device need be considered. For the three-lead device
shown, it must be the case that I 3 + I 4 + I 5 = 0 so one of the currents can always be
expressed using the sum of the other two. Three-lead devices will be seen later on.
Device leads are connected together to form circuits. A circuit will include one
or more paths through connected devices that allow a return to any starting point
without retracing steps. If there is no such path, then it is an “open circuit.” There
can be no (net) current in an open circuit. An open circuit corresponds to a dead-end
street. There is no net flow (over time) of cars into a dead-end street.
A very important consideration for circuit analysis is that any two (or more)
devices that have the same device rule, that is, the same relationship between
their currents and voltages, will behave exactly the same when placed in any
circuit. Such devices are electronic equivalents. Being able to replace a device with
another (possibly hypothetical) device will often prove convenient for solving and
understanding circuits.
A circuit is described on paper using a “schematic diagram” where each device
is represented by a symbol, referred to as a “schematic symbol.” Many of these

5Named after Georg Simon Ohm (1789–1854), a German mathematician and scientist. A Greek
omega is used since an “O” might be confused with a zero.
4 1 The Basics

Fig. 1.1 An illustration of two- and three-lead devices, and how the currents into and voltage
differences between the leads might be formally defined

Table 1.1 The simplest devices and their ideal behavior


Device Symbol Current–voltage relation
Wire, junction /\V = 0 between all leads, no matter what the current
of wires
Switch /\V = 0 if switch is closed, I = 0 if the switch is open

Resistor /\V = IR (“Ohm’s Law”), where R is a constant and is called


the “resistance.” Voltage decreases in the direction of current
flow
Battery (ideal /\V = V 0 , where V 0 is a constant called the “battery voltage.”
voltage source) Voltage increases from the “small” side to the “large” side

symbols are standardized, though some variation will be seen. Examples of symbols
used for the simplest ideal devices, previously introduced in introductory electricity
and magnetism, are shown in Table 1.1. A “node” is a connection point, a junction,
between two or more devices. In practice, nodes where three or more devices are
connected together will be most important.
Voltages are always measured as a voltage difference between two points in a
circuit. These voltage differences might be better described using a prefix delta (/\)
to indicate one is speaking of a difference, such as “/\V.” However, in electronics
the “/\” will routinely be omitted to simplify the notation. The current–voltage rela-
tionship for a device will, of course, always involve such a difference. For example,
Ohm’s law for resistors, which describes the current–voltage relationship for resis-
tors, is usually written simply as V = IR, rather than /\V = IR, even though the
latter might be clearer and/or more appropriate. When the voltage decreases across
a device in the direction of current flow, the change is referred to as a voltage drop.
The devices in Table 1.1 are all linear, meaning that if you change the voltage
(current) by any factor, the current (voltage) changes by the same factor. Circuits that
are constructed solely using linear devices are referred to as linear circuits. Linear
circuits can always be solved using linear mathematics. Additional linear devices
will be added in due course.
1.4 Resistors in Series 5

1.3 Kirchhoff’s Laws

Kirchhoff’s laws are obeyed by all circuits6 and provide the fundamental relations
used to solve for currents and voltages in a circuit.
• Kirchhoff’s Current Law (KCL): The sum of the currents entering any point in a
circuit equals the sum of the currents leaving that point. (That is, current is not
created or destroyed within the circuit—current is not “used up”.)
• Kirchhoff’s Voltage Law (KVL): The sum of the changes in the voltage, /\V i ,
around any closed path is zero. (A closed path is one that ends at the same position
where it starts.)
Example—Application of Kirchhoff’s Laws in a Simple Case
Apply Kirchhoff’s laws to the circuit illustrated by the schematic in Fig. 1.2 to get:
• KCL: The current through the 3 k resistor must be 2 mA (5 mA = 2 mA + 3 mA).
• KVL: V = 11 V because one must have V − (5 mA)(1 k) − (3 mA)(2 k) = 0.
(Here Ohm’s law, V = IR, was used to find the voltage change across the resistors.)

1.4 Resistors in Series

Two resistors are connected in series if they are connected so that the same current
must flow through both of them—there is no alternative current path.
Figure 1.3 shows several examples of resistors in series. Resistors in series form
a voltage divider. If the voltage across all of them is V 0 , then the voltage across the
resistors that are between two points A and B, designated V AB , will be given by the
“voltage divider equation7 ”.
E
R for all resistors between points A and B
V AB = V0 E . (1.1)
R for all the series resistors

This is a direct consequence of Ohm’s law and Kirchhoff’s voltage law.

6 As a fine point, Kirchhoff’s laws assume that the devices interact with each other only through
connecting wires. All of the basic physics inside the device is hidden. On the other hand, if two
devices interact to a significant degree via an electric or magnetic field, the rules may not apply as
written. In those cases, however, additional (hypothetical) devices can often be used to model the
effect of the interaction(s). Once those additional devices are included in the analysis, the rule is
again satisfied.
7 In this context, the upper-case Greek sigma, E, indicates a summation and is read as “the sum

of,” and the symbol R is a shorthand to represent “resistance values”.


6 1 The Basics

Fig. 1.2 A simple circuit to illustrate Kirchhoff’s laws

Fig. 1.3 Examples of resistors in series. In each case there is only one path for the current through
all the resistors

Fig. 1.4 Two examples of the use of the voltage divider equation for series resistors

Examples Use of the Voltage Divider Equation


• The voltage divider equation can be applied to find V AB for the circuit of Fig. 1.4a,
to get

R3 + R4
V AB = V0 . (1.2)
R1 + R2 + R3 + R4 + R5 + R6

• The voltage divider equation can be used to solve for V 1 for the circuit of Fig. 1.4b
to get

3.3 ku
V1 = 5V = 1.8V. (1.3)
1.2 + 3.3 + 4.7 ku

The voltage divider equation (e.g., in the examples above) is only valid if all the
resistors are in series. In practice, to measure the resulting voltage, a voltmeter must
be connected in parallel with a portion of the circuit. For example, to measure V 1
in the second example (Fig. 1.4b), the voltmeter would be connected across the 3.3
ku resistor—one lead from the voltmeter goes to each side of the resistor. If the
1.5 Resistors in Parallel 7

resistance of the voltmeter is too small, the current through the voltmeter will be
significant and that will affect the result—the voltmeter provides an alternate path
for some of the current so the resistors are no longer in series. The result will be
reasonably valid, however, if the current through the voltmeter is much less than the
current through the resistor. In this example, the voltmeter must have a resistance
very large compared to 3.3 ku for the reading to be accurate.
An “ideal voltmeter” has an infinite resistance (i.e., there will be no current through
the ideal meter). Most modern digital voltmeters will have an input resistance of 1
Mu or even larger, which is usually, but not always, large enough to be of little
consequence. Some older (analog) voltmeters may have a resistance small enough
so that the current through the meter cannot be routinely neglected.

1.5 Resistors in Parallel

Resistors are connected in parallel if the voltage across all of them must be the same
because they all are connected by (ideal) wires on both sides. Remember that there
is no voltage change across an ideal wire.
Figure 1.5 shows some examples of resistors in parallel. Parallel resistors form a
current divider. If the total current into the parallel combination is I 0 , then the current
through any given resistor, I k , can be computed using the current divider equation:

1/Rk
Ik = I0 E . (1.4)
1/R for all the parallel resistors

This is a direct consequence of Ohm’s law and Kirchhoff’s current law.


Examples Use of the Current Divider Equation
• Figure 1.6a illustrates the simplest case of two resistors in parallel. For that case,
the current divider equation yields

1/R1 R2
I1 = I0 = I0 . (1.5)
1/R1 + 1/R2 R1 + R2

Note the simple form of the current divider equation on the right when only
two resistors are involved—the current through one resistor is the total current
multiplied by the value of the other resistor, divided by the sum of the two. That

Fig. 1.5 Examples of


resistors in parallel. In each
case the voltage across all the
resistors must be the same
8 1 The Basics

Fig. 1.6 Two examples of the use of the current divider equation for parallel resistors

simple form is convenient to remember and use, but keep in mind that it only works
for pairs of resistors and does not generalize to three or more parallel resistors.
• For the portion of the circuit shown in Fig. 1.6b, the current I is given by

1/4.7
I = 3mA = 0.66mA. (1.6)
1/3.3 + 1/4.7 + 1/2.2

The current divider equation is only valid when all the resistors involved are in
parallel. To measure a current, you must insert a current measuring device in series
with some portion of the circuit—the circuit is broken and an ammeter is connected
across the break. In the second example above, the ammeter would be inserted in
series with the 4.7 ku resistor. If the resistance of the measuring device is significant
compared to the resistance in the circuit, the meter’s presence will affect the results.
In the second example, the resistance of the ammeter must be very small compared
to 4.7 k in order to get an accurate reading.
An “ideal ammeter” has no resistance (0 u). The internal resistance for many
ammeters will depend on the scale setting of the meter—that is, whether it is set to
measure larger or smaller currents.

1.6 Effective Resistance

If a circuit contains a number of resistors and some are in parallel and/or in series,
those combinations can be replaced with an “effective” or “equivalent” resistance,
often designated Reff or Req , without changing the behavior of the rest of the circuit.
It will be shown later that all combinations of resistors that ultimately connect to
two leads will have an equivalent, though it might not be simple to compute. The
equivalents for series and parallel resistors are easily derived using Kirchhoff’s laws:
E
• For series resistors: Re f f = REfor all the resistors
• For parallel resistors: 1/Re f f = 1/R for all the resistors.
Example—Equivalent Resistance—Parallel Resistors
The portion of a circuit in Fig. 1.7 has the three resistors are in parallel, so they can
be replaced with effective resistance:
1.6 Effective Resistance 9

Fig. 1.7 An example showing a group of parallel resistors that can be replaced with a single resistor
without affecting the rest of the circuit

1/Re f f = 1/3.3k + 1/4.7k + 1/2.2k = 0.97(1/k)


1
Re f f = k = 1.03k. (1.7)
0.97

Example—Equivalent Resistance—Combinations of Parallel and Series Resistors


Consider the portion of a circuit shown on the left in Fig. 1.8. As a first step, note
that all except the 15 k resistor are in series. We replace only those resistors with an
effective resistance:

Re f f = 3k + 4k + 6k + 5k = 18k. (1.8)

Once that is done, the 15 k resistor and the new effective resistance are in parallel.
They can be replaced, if desired, with a second effective resistance,
( )−1
1 1 15 · 18
Re' f f = + = k = 8.2k. (1.9)
15k 18k 15 + 18

That is, when placed in a circuit, the original combination of resistors will behave
the same as a single 8.2 k resistor, as shown on the right of Fig. 1.8.
Being able to replace series and parallel components with equivalents is one of
the most powerful ways to simplify circuits as part of an analysis. The process of
simplifying a circuit using equivalents is sometimes referred to as “circuit reduction”
and is discussed in more detail below. Additional circuit reduction techniques will
arise later.

Fig. 1.8 An example showing a portion of a circuit that can be reduced in two steps to a single
equivalent resistor
10 1 The Basics

1.6.1 Resistors in Parallel–Notation

Two resistors in parallel, R1 and R2 , can be replaced by an effective resistance, Reff ,


that is calculated using

R1 R2
1/Re f f = 1/R1 + 1/R2 → Re f f = . (1.10)
R1 + R2

For convenience, define an operator, ||, that acts on two variables and which produces
this result. That is,

R1 R2
R1 ||R2 ≡ . (1.11)
R1 + R2

This notation is used as a shorthand only, and be warned that the operator may not
behave like other mathematical operators, such as for multiplication and addition.
For example, the distributive law does not work, that is,

R1 ||(R2 + R3 ) /= R1 ||R2 + R1 ||R3 . (1.12)

Example—Describing Combinations Using the “||” Operator


Writing down the steps that would be used to compute an equivalent resistance, but
without solving at each step, is a way of describing the calculation. For example,
for the circuit of Fig. 1.9 the total effective resistance as seen by the battery can be
described symbolically as

Re f f = R1 + R3 ||(R2 + R4 ) + R5 , (1.13)

that you might read as


R-one in series with the combination of R-three in parallel with the series combination of
R-two and R-four, all of which is in series with R-five.

Can you provide a convincing argument to prove8 that for any three positive
values, A, B, and C,

A||(B||C) = (A||B)||C = (A||C)||B? (1.14)

That result means that the parenthesis can be left out without changing the result,
however remember that the || operator acts on only two values at a time.

8 Hint: this can be done rigorously without any algebra! Use some simple physics instead.
1.7 Solving Circuits with Circuit Reduction 11

Fig. 1.9 A circuit used to


illustrate the description of
the equivalent resistance
using the “||” operator

1.7 Solving Circuits with Circuit Reduction

As mentioned above, many circuits can be solved using a process known as circuit
reduction—a more complicated circuit is reduced to something simpler using an
electronic equivalent. The effective resistance of series and parallel resistors can be
most useful for this purpose.
Example—Circuit Reduction Using Parallel and Series Resistors
Determine the currents through all of the resistors for the circuit in Fig. 1.10a.
Recall:
• Resistors are in series if the current through them MUST be the same (because
there is no other path). In the circuit of Fig. 1.10a there are NO series resistors.
• Resistors are in parallel if the voltage across them MUST be the same (i.e., both
ends are connected by wires.)
In the circuit of Fig. 1.10a, the 800 u and 1.5 ku resistors are in parallel (and no
others). As far as the rest of the circuit is concerned, these two can be replaced with
a single resistor with a value

800 · 1500
R1 = 800u||1.5ku = u = 522u, (1.15)
800 + 1500

and so now the circuit is as shown in Fig. 1.10b.


Examining this new circuit, there are no resistors in parallel, but the 470 and 522
u resistors are in series. Hence, as far as the rest of the circuit is concerned, they can
be replaced with a single resistor

Fig. 1.10 Using equivalents for series and parallel resistors, the circuit shown in (a) is reduced in
several steps in several steps, (b)–(d), to the simple circuit shown in (e)
12 1 The Basics

R2 = 470u + 522u = 992u, (1.16)

and now the circuit is as shown in Fig. 1.10c.


Examining Fig. 1.10c, there are no resistors in series, but the 1 k and 992 u
resistors are in parallel. Hence, as far of the rest of the circuit is concerned, they can
be replaced with a single resistor with a value

1000 · 992
R3 = 1ku||992u = u = 498u, (1.17)
1000 + 992

and the circuit now is that of Fig. 1.10d.


Examining Fig. 1.10d, there are no parallel resistors, but the 500 and 498 u
resistors are in series. Hence, as far as the rest of the circuit is concerned (which now
is only the battery and wires) they can be replaced with a single resistor

R4 = 500u + 498u = 998u. (1.18)

Now the circuit is the simple circuit of Fig. 1.10e. The current delivered by the
9 V battery to the rest of the circuit (which, to the battery, looks like a single 998u
resistor) is easily computed, using Ohm’s law, to be

9V
I1 = = 9.0mA . (1.19)
998u
At this point all of the original circuit has been lost except for the battery. Working
backward, as needed, the currents through (and hence the voltages across) all of the
original resistors can be determined. Refer to Fig. 1.11.
Going back one step, whatever current goes through the 998 u equivalent resistor
is really going through the 500 and 498 u resistors which are in series (and hence
they must have the same current).
Now the 498 u resistor came from two resistors in parallel. Those two must have
the same voltage across them and that voltage is the same as the voltage across the
498 u resistor. That is, V = 9.0 mA · 498 u = 4.48 V, so find the currents that give

Fig. 1.11 Starting with the result from Fig. 1.10e, the currents through each resistor are determined
by working in reverse
1.8 Solving Circuits with Algebra 13

this voltage for the two parallel resistors, or use the current divider equation, which
is equivalent, to get the currents shown in Fig. 1.11b.
Now the 992 u resistor came from two resistors in series, each of which must have
the same current (4.52 mA, Fig. 1.11c). The 522 u resistor is really two resistors
in parallel, so the voltage across both of them is V = 4.52 mA · 522 u = 2.36 V.
Once again, find the appropriate currents that give you this voltage for each of the
two resistors to get the currents shown in Fig. 1.11d. Now “everything” is known
and the circuit has been solved.
Depending upon what question is to be answered, it may not be necessary to
proceed this far. For example, if only the current delivered by the battery was required,
the process could have stopped several steps ago. If the voltage across the 800 u
resistor was desired, it was available at the previous step (it’s 2.36 V).

1.8 Solving Circuits with Algebra

Not all circuits contain components that can be easily replaced with an equivalent. In
those cases, Kirchhoff’s laws and algebra can be used. Several different approaches
are useful in order to generate the equations to be solved.

1.8.1 Branch and Mesh Currents

There are many ways to solve circuits algebraically using the equations that result
from Kirchhoff’s laws. Two common procedures to set up those equations are the
branch and mesh methods. A third method, known as nodal analysis, will be shown
separately. All of these methods start with Kirchhoff’s laws to generate a number of
equations. For linear circuits, those equations will be mathematically linear and can
always be solved, at least in principle.
• Branch Method:
A current is defined for each portion of the circuit that may have a different current.
That is, between all adjacent pairs of nodes where three or more devices connect.9
• Mesh (or Loop) Method:
Each possible loop in the circuit is given a label, up to the number of independent
loops necessary to include each component at least once. The advantage is that
Kirchhoff’s current law (KCL) will be automatically satisfied and there are fewer
equations to solve. The disadvantage is that the current in any particular portion

9 Defining the current means giving it a name and defining which direction is to be considered
positive. If a negative value for the current results, that simply means the arrow was drawn opposite
to what really occurs. This process is analogous to defining the positive the x- and y-axes for a
mechanics problem.
14 1 The Basics

Fig. 1.12 The circuit shown in (a) is to be solved. For the branch and mesh methods, the first step
is to define the currents such as is shown in parts (b) and (c) respectively

of the circuit may be the sum of several such mesh currents, which may be an
inconvenience.
A simple circuit is illustrated in Fig. 1.12a. The description of the circuit using
branch currents is shown in Fig. 1.12b, and using mesh currents in Fig. 1.12c. Of
course, what actually happens in the circuit must not depend on the description.
Hence, it is always the case that one description can be converted to the other by
equating the currents through components. In this example, it must be that

Ia = I p , I c = Iq , I b = I p − Iq . (1.20)

Note that the variable names you use are up to you. It is very common to use the
same letter (e.g. “I”) with different subscripts (whether you use number, letters, or
signs of the Zodiac as subscripts, is up to you).10 As a further note, in electronics the
symbol “i” (lower case) is sometimes used for currents. In this text the lower case
“i” will be used for another purpose and so its use as a label for currents is avoided.

1.8.2 Example—Using Kirchhoff’s Laws

As an example, the currents in a circuit will be found with Kirchhoff’s laws using
both branch currents and mesh currents. Of course, the actual current found in any
part of the circuit must be the same for the two solutions. For many circuits, the mesh
method is more efficient and therefore is preferred. If the correctness of the answer
is particularly important, solve using both methods and compare the results.

10 What is now simply called “current” used to be called “current intensity,” or in French, the
language of Ampère, intensité du courant. Hence the use of “I” to represent that “intensity.”
1.8 Solving Circuits with Algebra 15

Goal: Find all the currents for the circuit of Fig. 1.12.
Using branch currents:
(1) Define the currents in each branch (i.e. give them a name and direction). This
is illustrated in Fig. 1.12b.
(2) Write down KCL equations at nodes (junctions). Remove any duplicate or redun-
dant equations. KCL is trivially satisfied elsewhere. In this case there is one
unique KCL equation,

Ia = Ib + Ic . (1.21)

(3) Write down KVL for enough loops to include all components at least once,
adding the voltage changes along the way. The loops can be in either direction.
There are three (simple) loops here, generating the following equations:

V0 − Ia R1 − Ib R3 − Ia R5 = 0
V0 − Ia R1 − Ic R2 − Ic R4 − Ia R5 = 0 (1.22)
Ib R3 − Ic R2 − Ic R4 = 0.

There are three unknown currents so three independent equations will be neces-
sary. The KCL equations are (almost) always required. Here there is one unique
KCL equation. Hence only two of these KVL equations are needed. Note, for
example, in this case the third KVL equation can be generated from the first two
(by subtracting the first from the second), hence, given any two of these KVL
equations, the third will provide no additional information.
(4) Now solve the equations for the desired quantity(ies).
As an example, use the first and third KVL equations. The third equation can
be rearranged to get

R3
Ic = Ib , (1.23)
R2 + R4

and putting this into the KCL equation gives


( ) ( )
R3 R2 + R4 + R3
Ia = Ib 1 + = Ib
R2 + R4 R2 + R4
( )
R2 + R4
Ib = Ia , (1.24)
R2 + R3 + R4

that can be substituted into the first KVL equation to get


( )−1
R3 (R2 + R4 )
Ia = V0 R1 + + R5 , (1.25)
R2 + R3 + R4
16 1 The Basics

and this value can be substituted back into the previous equation relating I a and
I b to find I b , and then that value is used to find I c .
Using mesh currents:
(1) Define mesh currents (each current goes around a complete loop, and the total
current at any location may be the sum of several mesh currents). Each compo-
nent must have at least one mesh current through it. There are three simple
loops here: smaller loops on the left and right, as in Fig. 1.12c, and a larger
loop around the outside (not shown). Very complicated loops, such as a figure-
eight or involving multiple passes through a device, are also possible, but not
recommended. Keep it simple.
(2) Since mesh currents automatically satisfy KCL, there are no KCL equations to
write down.
(3) Write down KVL equations—at a minimum, there should be enough to include
all components at least once. The loops used for these equations do not need to
be the same as those used to define the currents. Note that for this example, the
current through R3 involves the sum of two mesh currents with opposite signs.
Three equations generated from the three simple loops are
( )
V0 − I p R1 − I p − Iq R3 − I p R5 = 0
V0 (− I p R1 −
) Iq R 2 − Iq R 4 − I p R 5 = 0
(1.26)
I p − Iq R3 − Iq R2 − Iq R4 = 0.

There are two unknown currents and so only two of these equations will be
necessary. Rewriting the second and third,

I p (R1 + R5 ) + Iq (R2 + R4 ) = V0
(1.27)
Iq = R2 +RR33 +R4 I p ,

and putting the second of these two results into the first yields
( )
(R2 +R4 )
I p R1 + R5 + RR23 +R 3 +R4
= V0
( )−1 (1.28)
(R2 +R4 )
→ I p = V0 R1 + R5 + RR23 +R 3 +R4
,

which can then be substituted back into the equation just above get I q . Note the
connection between this solution and the solution using branch circuits. Here
I a = I p , I c = I q , and I b = I p – I q , as they should.
For both of these algebraic methods it is easy to generate more equations than
necessary. Such a situation can be easily resolved. It is also possible to generate
too few independent equations. If the equations are not solvable or reduce to trivial
results (such as 0 = 0) then go back and try a different set of equations.
1.8 Solving Circuits with Algebra 17

1.8.3 Nodal Analysis

An additional way to analyze a circuit is known as nodal analysis. Nodal analysis


may be somewhat inefficient when applied to an entire circuit, but it may be very
useful for smaller circuits or for the analysis of a small portion of a circuit. Nodal
analysis also works well for automated (i.e., computerized) solutions.
To apply nodal analysis, the voltage (relative to some convenient common refer-
ence) at each unique node is given a label. Kirchhoff’s voltage law then allows you to
find the current between each pair of nodes in terms of those voltages and the device
rules (e.g., Ohm’s law). Kirchhoff’s current law is then used at each node, writing
the currents in terms of the node voltages. If you have N nodes, there will be N − 1
unique equations that can then be solved. Again, in the general case all the devices
must be included at least once to get a complete solution.
Example—A Solution Using Nodal Analysis
Start by labeling the nodes and defining the currents, such as is illustrated in Fig. 1.13.
In this example there are two nodes, one of which can be taken to be the voltage
reference. Kirchhoff’s current law gives, as previously seen above for the analysis
using branch currents,

Ia − Ib − Ic = 0, (1.29)

where
V2 − V1 + V0 V1 − V2 V1 − V2
Ia = ; Ib = ; Ic = . (1.30)
R1 + R5 R3 R2 + R4

If the lower node is taken as the reference, equivalent to using V 2 = 0 V, then


KCL gives
( ) ( )
1 −1 1 1
V0 + V1 + + = 0, (1.31)
R1 + R5 R1 + R5 R3 R2 + R4

Fig. 1.13 The first step when solving the circuit of Fig. 1.12a using nodal analysis is to define a
variable associated with each of the branch currents and the node voltages
18 1 The Basics

where the only unknown is V 1 . It is straightforward to solve for V 1 , and that value is
used to solve for the currents. Of course, the resulting currents through the devices
must agree with those found above using the other methods.
With nodal analysis, there may be a larger number of equations to solve, but many
of them will often be, in a practical sense, simpler than what is encountered when
using branch or mesh currents.

1.9 The Ideal Current Source

Another basic linear device seen in electronics is the “ideal current source.” The
ideal current source is a device that provides a constant current output regardless of
the voltage across it. That is, I = I 0 , and V is whatever is necessary to maintain that
current. Schematically the current source appears as shown in Fig. 1.14, where the
arrow indicates the direction for positive current (the “I” may or may not be present).
Example—Simple Example Involving a Current Source
In the circuit of Fig. 1.15a, what is the voltage across the current source?
Using circuit reduction, the resistors can be replaced by an equivalent with a value
1k + (2k||3k) = (1 + 6/5)k = 2.2k. The equivalent circuit as seen by the current
source is shown in Fig. 1.15b.
The current I is 30 mA and so, using Ohm’s law, the voltage across the equivalent
resistor is 2.2 k · 30 mA = 66 V. In this case KVL requires that that must also be the
magnitude of the voltage across the current source.

Fig. 1.14 The schematic symbol for a constant current source is shown in (a). Some alternate
symbols are sometimes used for current sources, such as those in (b). The diamond shaped symbol
is often used to indicate a “dependent source,” where the current depends on what is happening
somewhere else in the circuit

Fig. 1.15 One method to find the voltage across the current source for the circuit shown in (a), is
to use circuit reduction to find an equivalent resistance, as shown in (b)
1.9 The Ideal Current Source 19

Most commercial power supplies can be modeled as constant voltage sources,


though there are exceptions and constant current supplies are certainly available.
Constant current sources will be most prominent later, principally as part of a
simplified model for a transistor.
Example—Nodal analysis with both a voltage and current source
The goal is to solve for the currents in the circuit of Fig. 1.16. There are three
node voltages labeled, V a , V b , and V c , and five branch currents. Four of the branch
currents are labeled with variables I 1 to I 4 . The fifth branch current is that of the
current source, which has a known value (2 mA).
Kirchhoff’s current law (KCL) will need to be applied at each of the nodes,
generating the following equations:

I1 = I2 + I3 + 2mA
I3 + I4 = I1 (1.32)
I4 = I2 + 2mA.

Now look at the voltage changes along the paths between adjacent nodes to
compute the currents:

I3 = (Va − Vc )/3k = 3V/3k = 1mA


I2 = (Va − Vb )/2k
I4 = (Vb − Vc )/1k. (1.33)

There are three node voltages, any one of which could be taken to be the refer-
ence—remember that only voltage differences will matter when determining the
currents. For the sake of this example, use V c as the reference. That is equivalent to
taking V c = 0 V (since the difference between V c and itself is 0 V, why not just call
it 0 V?).
Note that there are 6 unknowns: four currents and two voltages. Thus, six inde-
pendent equations are required. (If the mesh current method above was used, there
would have been only 3 unknowns.)
Solving this circuit yields:

Fig. 1.16 A circuit


involving both a current
source and a voltage source
that is to be solved using
nodal analysis
20 1 The Basics

Va = 3V
I3 = 1mA
1V
3V = (2k)I3 + (1k)I4 = (2k)I3 + (1k)(I2 + 2mA) → I2 = = 0.33mA
3k
I4 = 2.33mA
Vb = 2.33V
I1 = 2.33mA + 1mA = 3.33mA (1.34)

and the voltage across the current source is 0.666 V.

1.10 The Ground and Common Connections

The schematic representation of the “ground” and “common” connections are shown
in Fig. 1.17. When these appear in circuit diagrams, all of the “common” connection
points are to be connected together with a wire (or equivalent). The “ground” is a
special type of common connection that includes a wire which is actually connected
to the Earth. Sometimes such a connection is called “earth” instead of the “ground”
connection. Inside a building, this connection to the Earth might not be obvious and
there may be a long route to get there.
The Earth is very large, is able to conduct electricity, and is presumed to be a
good source and sink for excess charge. When you ground (or “earth”) a circuit you
are connecting it to the Earth. Such a connection is presumed to provide a constant
electric potential, usually taken to be 0 V. The person operating an electronic device
is typically in contact with the Earth, either directly or indirectly, so there will be no
voltage difference between the operator and such a ground connection. Hence such
a location will not pose a shock hazard.
Sometimes the ground connection will be used in a schematic diagram when there
is no actual connection to the Earth. What is usually meant is a common connection
and/or a connection to a chassis or other larger metal object which acts like an
(idealized) earth connection, at least for the intended use of the circuit. There is
some sloppiness in the use of the ground symbol.
For the sake of circuit analysis, the three schematics of Fig. 1.18 would be solved
in exactly the same way.
Some complicated circuits may have more than one common connection. In such
cases a label will appear inside or next to the symbol. All common connections
with the same label are connected together, but those with different labels are not
connected to each other.

Fig. 1.17 Schematic


symbols for ground and
common connections
1.11 Multiple Sources—The Superposition Theorem 21

Fig. 1.18 Since all common connections, shown in (b), are connected to each other and all ground
connections, shown in (c) are connected to each other, the analysis for all three of these circuits will
be identical

1.11 Multiple Sources—The Superposition Theorem

The superposition theorem can be used to solve circuits containing multiple sources
using any of the methods described above. Superposition will also be very important
for the treatment of time-dependent voltages and currents.
The superposition theorem says that for any linear circuit containing more than
one independent source, the circuit can be solved by considering one source at a
time, with all the other sources “turned off,” and then adding those results together.
That is, the solution for the sum is the sum of the solutions. This is a special case of
a more general result from basic electricity and magnetism (E&M) and, even more
generally, linear mathematics.
In this context:
• A voltage source that is “turned off” is a voltage source fixed at 0 V—such a
source is equivalent to a wire.
• A current source that is “turned off” is a current source fixed at 0 A—such a source
is equivalent to an open circuit (no connection).
Example—Use of the Superposition Principle
Determine the current, I, delivered by the 9 V battery in the circuit of Fig. 1.19a. The
current delivered by the 9 V battery is the same as the current through the 1 k resistor.
First, look at that current due to the 9 V battery when the 3 V battery is “off” (i.e.,
replaced with a wire, Fig. 1.19b). Then I 1 = 9 V/Reff where Reff = 1 k + 2 k||3 k =
2.2 k, so I 1 = 4.09 mA.
A common mistake is to claim that I 1 is the current delivered by the 9 V battery.
This is not true. The analysis is not yet complete.

Fig. 1.19 The circuit shown in (a), that contains two sources, is to be solved using superposition
by considering each source separately with the other “off,” as shown in (b) and (c)
22 1 The Basics

Now look at the current due to the 3 V battery, with the 9 V battery “off”
(Fig. 1.19c). Then I 2 = I ' 3 k/(1 k + 3 k) = 3I ' /4 (using the current divider), and I '
= 3 V/Reff where Reff = 2 k + 3 k||1 k = 2.75 k. Hence, I ' = 1.09 mA and so I 2 =
0.82 mA.
To get the result for the original circuit, add these together taking direction into
account. Since the directions used for I 1 and I 2 were opposite, and I 1 is in the
direction of the current desired (I), we subtract I 2 from I 1 .
Hence, I = (4.09 − 0.82) mA = 3.3 mA is the current delivered by the 9 V battery.
As an exercise, compute the current through the 3 k resistor and the current
delivered by the 3 V source for the same circuit.11

1.12 Electrical Power

Power is the rate of change of energy and is measured in watts (W),12 where 1 W
= 1 J/s. In mechanics, power, P, is the dot product of force and velocity. Power is
supplied to an object if the force on the object and the velocity of the object are in the
same direction. In electronics, the power supplied to a device, or supplied by a device,
is given by the product of the voltage across the device and the current through the
device, P = VI. The power is being supplied by the device if the current through
the device goes from a lower voltage to a higher voltage. Power is being absorbed
by the device (e.g., turned into another form of energy, such as heat, chemical, or
mechanical energy) if the current flows through the device from a higher voltage to
a lower voltage.
For the devices shown in Fig. 1.20 (and assuming I is positive), power is being
supplied to the device on the left, while power is being supplied by the device on the
right. For a resistor, Ohm’s law can be used to see that power, P, is always supplied
to a resistor, and never by the resistor, and P = VI = I 2 R = V 2 /R.
The watt (W) is an SI unit. Equivalent units are: 1 W = 1 V · A = 1 A2 · u = 1
V /u = 1 J/s = 1 kg · m2 /s2 .
2

Note that since power involves the product of voltage and current, superposition
does not apply directly to power. That is, the power supplied when there is more than
one power source is not the sum of the power from each source considered separately.

11 Answers: 1.9 mA down through the 3 k resistor, and −1.4 mA delivered by the 3 V battery—that
is 1.4 mA is going into the 3 V battery.
12 Named for the Scottish inventor and engineer, James Watt (1736–1819).
1.13 Additional Application—The Kelvin-Varley Divider 23

Fig. 1.20 If the current goes from higher to lower voltage, shown on the left, power is being
supplied to the device. If the current goes from lower to higher, shown on the right, power is being
supplied by the device

1.13 Additional Application—The Kelvin-Varley Divider

A potentiometer is a variable resistor with three leads. Hence, the potentiometer can
be used as a variable voltage divider. In the schematic (Fig. 1.21), the connection with
the arrow can move and is called the “wiper.” The total resistance between the outer
two leads is fixed and is cited as the value, R, of the potentiometer. The resistance
from the wiper to the two outer contacts (R1 and R2 ) varies from 0 to the full value
R, and no matter what the position, it is always the case that R1 + R2 = R. In simple
potentiometers the wiper makes a spring-loaded contact with a resistive material and
a continuous range of values is obtained. The potentiometer can also be used as a
variable resistor (a “rheostat”) if the center lead and only one of the outer leads is
used.
The wiper is often moved using a dial or a slider. The two common types of such
devices will use a “linear” or “audio” taper. That is, the relationship between R1
and the dial or slider position is either linear or non-linear (in a manner which is
approximately exponential, and is called an “audio” or “log taper”). One reason the
latter is convenient is because our ears perceive volume approximately linearly with
the logarithm of output amplitude, and the log is the inverse of the exponential—our
hearing straightens out the exponential changes in amplitude.
While potentiometers are often used as simple variable voltage dividers, a preci-
sion voltage divider can be used to compare two voltages (or “potentials”) very
precisely. Connected as shown in Fig. 1.22, the divider is adjusted for a null
reading giving the ratio of the known reference to an unknown being measured.

Fig. 1.21 The schematic symbol for a potentiometer shows three leads and is an adjustable voltage
divider. The resistance between the outer two leads is fixed. The resistance between the center lead
and each outer lead is adjusted, typically by turning a knob
24 1 The Basics

Fig. 1.22 A potentiometer can be used with a known reference voltage to determine an unknown
voltage to high accuracy. A calibrated potentiometer is adjusted until the null detector (e.g., an
ammeter) reads zero

Fig. 1.23 A schematic showing how a 10 k Kelvin-Varley potentiometer is constructed

When adjusted for a null, no current, and hence no power, will be drawn from the
unknown source, and hence the reference and potentiometer act as if they have infinite
resistance to the unknown.
The Kelvin-Varley voltage divider is a potentiometer design that allows for great
precision and a discrete readout. The potentiometer is made from discrete, matched,
precision resistors with a schematic such as is shown in Fig. 1.23. In each of the
“digital” sections there are eleven resistors. Two of those eleven will be in parallel
with a resistance equal to those two in series. As you turn each dial, two contacts
move together to adjacent positions along the voltage divider. The last stage is usually
a continuous reading dial.
1.14 Problems 25

For the switch positions shown in the schematic, V out = 0.164xx V in , where the
last digits (‘xx’) depend on the position of the wiper on the 80 u potentiometer. Here
the effective resistance seen at the input (if nothing is connected on the output) is
10 k no matter what the dial settings.
Note that when the null is achieved for the comparison circuit above, there will
be no current into or out of the final output of the divider and so the output voltage
(V out ) can be determined as if nothing were connected and accuracy is maintained.
For accurate results in other cases, the output of the divider must be attached to a
device or circuit that draws only a negligible current.
To analyze this Kelvin-Varley potentiometer in the case where there is no output
current, start from the left and note that between the two points connected by any
pair of switches, half of the current proceeds down and half goes through the switch
connections to the right. For example, for the circuit of Fig. 1.23, if I 200 is the current
entering the string of 200 u resistors, then the current going to the 40 u resistors is
I 200 /2.

1.14 Problems

1. (a) For the circuits shown in Fig. 1.P1, find the current through each of the resis-
tors using Kirchhoff’s current and voltage laws to set up equations, then solve
the equations (do not use parallel/series resistor equivalents, circuit reduction,
etc.). (b) Using your results from (a), find the voltage across the 3 k resistor for
each circuit.

Fig. 1.P1 Problem 1

2. For the circuits of Fig. 1.P2, use circuit reduction and superposition to determine
the current through each of the resistors. What is the power delivered by each
of the sources?
26 1 The Basics

Fig. 1.P2 Problem 2

3. For the circuit of Fig. 1.P3, the current through the 1 k resistor is 1.0 mA. What
is the battery voltage, V?

Fig. 1.P3 Problem 3

4. Use Kirchhoff’s and Ohm’s laws to derive the (a) voltage divider, (b) current
divider, and effective resistance results for (c) series and (d) parallel resistors.
5. A real voltmeter that acts as a 1 Mu resistance is used to measure the voltage
across a resistor, as shown in the circuits of Fig. 1.P5. For each circuit, estimate
the difference between the measured voltage and the voltage that would be
measured with an ideal voltmeter.

Fig. 1.P5 Problem 5

6. For the circuit shown in Fig. 1.P6, (a) what is the current through R0 and (b)
show that when R0 gets large, in particular when R0 >> R1 + R2 , that the voltage
across the current source is close to V − I 0 (R1 + R2 ), independent of the value
1.14 Problems 27

of R0 , and that if it is also the case that R0 >> V /I 0 , then I 1 ≈ I 0 . That is, if these
two conditions are satisfied, the circuit acts the same as if R0 were not present.

Fig. 1.P6 Problem 6

7. What is the voltage across the 3 k resistor in the circuit of Fig. 1.P7?

Fig. 1.P7 Problem 7

8. The circuit of Fig. 1.P8 consists of a resistor “tree” having four levels connected
in series and powered by a battery with voltage, V 0 . The uppermost level has
one resistor and each lower level has twice as many resistors as the level above
it. If all of the resistors have the same value, R, what is the current from the
battery for this four-level tree? What is the current in the limit that there are an
infinite number of such levels?
9. Using “ = ” and “ > ”, rank order the magnitudes of the five branch currents
for the circuit of Fig. 1.P9 from largest to smallest, without first solving for the
currents.
10. Resistors are sometimes used to model the electrical behavior of incandescent
light bulbs. What resistance value should be used for a 50 W bulb designed to
work with (a) a 12 V supply, (b) a 120 V supply, and (c) a 240 V supply.
11. Show that (a) the effective resistance of a series combination of resistors is
always larger than the largest individual resistance, and (b) that the effec-
tive resistance of a parallel combination of resistors is always smaller than
the smallest individual resistance.
28 1 The Basics

Fig. 1.P8 Problem 8

Fig. 1.P9 Problem 9


Chapter 2
Additional Theorems

Kirchhoff’s laws, parallel and series resistor equivalents, along with the superposition
principle, will go a long way toward determining the voltage-current relationships
in many simple circuits involving resistors and ideal sources. Later, these staples of
circuit analysis will be applied more generally to include other linear circuit elements.
However, before adding additional circuit elements, a few additional theorems, and
examples of their use, are presented.

2.1 Thevenin and Norton Equivalents

Consider a box with two external leads. Inside the box is any number of resistors
and linear sources (voltage sources and/or current sources) connected in any way
(for example, see Fig. 2.1a). When this box is connected into a circuit, and assuming
appropriate values are chosen for V th , Rth , I n , and Rn , the circuit will behave the same
if either the Thevenin or Norton equivalent is used instead (Fig. 2.1b, c).
The Thevenin and Norton equivalents are used, for example, to analyze or better
understand a circuit by reduction. That is, even a very complicated piece of a circuit
can be replaced by a very simple equivalent, simplifying the rest of the analysis. The
problem is to find appropriate values for the parameters V th , Rth , I n , and Rn .
It is important to realize that once a complicated circuit has been replaced with
one of these equivalent circuits, all details of what might be going on inside the “box”
that was replaced are hidden from view.

2.1.1 Determining the Thevenin and/or Norton Parameters

The appropriate values needed for the equivalents can often be determined by consid-
ering limiting cases. The voltage-current relationship for both the Thevenin and

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 29


B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_2
30 2 Additional Theorems

Fig. 2.1 Any configuration of resistors and ideal sources, such as illustrated in (a), can be simplified
to the (b) Thevenin equivalent or (c) Norton equivalent circuit if appropriate component values are
used

Norton equivalents can be described by a single straight-line plot of current as a


function of voltage, such as shown in Fig. 2.2. Thus, the behavior of the entire
complicated “box” of components is also described graphically by the same straight
line. That line can be determined by finding any two distinct points on the line. On
paper, at least, the intercepts with the axes form a convenient pair of points to use.
Hence, the values for the equivalent circuits can be found by noting the following:
• The intercept with the V-axis is the voltage between the two wires from the box
when the box is not connected to anything. If it is not connected, the net current
out of the box is zero. This intercept is the “open circuit voltage.” Comparing the
original circuit to the equivalent circuits, in each case with nothing connected,
that open circuit voltage is the same as the Thevenin equivalent voltage, V th .
• The intercept with the I-axis is the current through the two leads of the box when
the output is short-circuited—that is, an ideal wire is attached between them

Fig. 2.2 The Thevenin and


Norton equivalents, and thus
the more complicated circuit
that they replace, will have a
straight-line relationship
between current and voltage.
The intercepts between that
line and the axes are
convenient values to use to
determine V th and I N
2.1 Thevenin and Norton Equivalents 31

Fig. 2.3 The circuit shown in (a) is replaced by its Thevenin equivalent. Intermediate steps are
shown in (b) and (c) and the result in (d)

forcing the voltage to zero.1 That current is the “short circuit current” and is equal
to the Norton equivalent current, I n .
• The equivalent resistance is given by the negative inverse slope of the line, equal
to V th /I n . Note that it is always the case that for a given circuit, Rth = Rn .
In general, any two points on the line can be used—they do not need to be the
intercepts. The intercepts are often algebraically convenient to find, and from them
the slope. The Norton equivalent can always be found from the Thevenin equivalent,
and vice versa, using the relations Rth = Rn and I n Rn = V th . Sometimes it is easier
to find one or the other equivalent and then convert as needed.
An alternate (and often easier) way to determine Rth = Rn is to find the equivalent
resistance of the components in the box when all the sources within it are adjusted so
their outputs are zero. That is, they are “turned off.” This method is directly related to
the superposition theorem seen in Chap. 1. Remember that in this context a voltage
source that is “off” (0 V) behaves the same as a wire, and a current source that is
“off” (0 A) behaves the same as an open circuit. While it might not be possible to
turn off the sources for a real box, it can always be accomplished on paper as part of
circuit analysis.
Example—Finding Thevenin (and Norton) Equivalents
As an example, find the Thevenin equivalent for the portion of a circuit shown in
Fig. 2.3a.
The Thevenin equivalent voltage equals the voltage between the two free wires
when nothing is attached to them. With nothing attached, there will be no current
through the 2 k resistors so the change in voltage across those resistors is zero (from
Ohm’s law). Hence, start by looking at the simpler circuit of Fig. 2.3b, which will
have the same V th . As long as nothing is connected to the wires on the right, the two
4 k resistors and the 5 k resistor are in series—whatever current goes through one
must go through the others—and so the voltage across the 5 k resistor, which is equal
to V th , can be found using the voltage divider equation,

5k 25
Vth = 5V = V = 1.92V. (2.1)
4k + 5k + 4k 13

1 Note that you can think about these limiting cases on paper, but may not be able to realize them
in practice without destroying some circuit components.
32 2 Additional Theorems

To find Rth , return to the original circuit but turn the voltage source off (i.e., replace
it with a wire) and find the equivalent resistance as seen by anything connected to this
circuit, that is, as seen from the right as shown in Fig. 2.3c. Note that the simplification
made above to find V th (i.e., ignoring the 2 k resistors) obviously does not apply when
finding Rth . Hence,

Rth = 2k + 5k||(4k + 4k) + 2k = 5k||8k + 4k = (5 · 8)k/(5 + 8) + 4k = 7.08k.


(2.2)

Hence, the original piece of the circuit in Fig. 2.3a can be replaced with the simpler
equivalent piece shown in Fig. 2.3d. Such a replacement will not change the behavior
of the rest of the circuit. The Norton equivalent current will be the current when the
outputs of the Thevenin equivalent are short-circuited. In this case, I n = 1.92 V/
7.08 ku = 0.271 mA.

2.1.2 How Is This Used for Circuit Reduction?

Suppose the circuit in Fig. 2.4a is to be analyzed to determine the current through
the 1 ku resistor. Then, using the analysis from the previous example, everything
except the 1 k resistor can be replaced by its Thevenin equivalent, resulting in the
circuit of Fig. 2.4b. That circuit is easily solved to give a current I = 1.92 V/(8.08
ku) = 0.24 mA.
A common mistake at this point is to say that the current from the original 5 V
battery is also 0.24 mA. This, of course, is not correct. The 5 V battery also sends
current through the 5 k resistor. After the Thevenin equivalent is created, the current
through the 5 k resistor is no longer visible in the diagram—that current is hidden,
as is the original 5 V battery.2

Fig. 2.4 To determine the


current through the 1 k
resistor, the rest of the circuit
can be replaced with its
Thevenin equivalent

2For this particular circuit, the actual current delivered by the 5 V battery is 0.48 mA, exactly twice
what flows through the 1 k resistor. Can you see why it is exactly twice for this example?
2.1 Thevenin and Norton Equivalents 33

2.1.3 Equivalent for an Infinite Array of Resistors

Thevenin’s (or Norton’s) equivalent for any number of resistors connected in any
way, and in the absence of voltage or current sources, will be a single equivalent
resistor, Req . In addition, there are times when an array of components, considered in
the limit that their number becomes infinite, can be a useful model for a real system.
Examples include models for certain transmission lines (long cables). With that in
mind, consider the infinite array of resistors of Fig. 2.5a.
The trick for solving many problems is to exploit a symmetry of the system. For
this infinite array the basic symmetry is translational. If the first pair of resistors is
removed (or the first n pairs), the remaining array is still infinite and looks exactly the
same as the original. Therefore, the remaining array can be replaced with Req , which
is not yet known, but is presumed to exist. Then the circuit looks like Fig. 2.5b.
Using circuit reduction,
( )
R1 R2 + Req R2 Req
Req = R1 + R2 ||Req = , (2.3)
R2 + Req R2 + Req

and then multiplying to get rid of the denominator on the right and rearranging results
in the quadratic
2
Req − R1 Req − R1 R2 = 0. (2.4)

Using the quadratic equation, the solutions are


/
R1 ± R12 + 4R1 R2
Req = , (2.5)
2
where clearly one must choose the plus sign (or else Req is negative, which means it
supplies energy rather than dissipates energy).
One special case is when R1 = R2 in which case

Fig. 2.5 The equivalent resistance of the infinite array of resistors shown in (a) can be deter-
mined using the translational symmetry (b), which is then solved using parallel and series resistor
substitutions
34 2 Additional Theorems
() √ (
1+ 5
Req = R1 . (2.6)
2

The ratio on the right is known as the “Golden Ratio” which is claimed to have
some aesthetic appeal in the Arts. Apparently, it can show up in electronics as well.
Solutions for some two-dimensional infinite resistor arrays are also possible. Some
such solutions are presented by [1].

2.2 The Wheatstone Bridge

The Wheatstone bridge circuit provides a means to match resistance values to very
high precision.3 The bridge circuit when used to make such a measurement is often
drawn as shown in Fig. 2.6a. The circuit element “M” is a meter. Using Thevenin’s
theorem, the meter can be replaced with a resistor (this assumes the meter does
not also act as a source). The goal is to predict the reading on the meter. To solve
the circuit using Kirchhoff’s laws and equations, the setup as shown in Fig. 2.6b
could be used. With loop currents, three loops are required and three equations will
result. It is possible, though a bit cumbersome, to produce values using those three
equations. Such a solution can be found in many electronics texts. It is much simpler
to understand and produce a result using the theorems that have been presented above.
To solve for the voltage across, and current through, the detector, R5 , it is straight-
forward to show that the rest of the circuit can be reduced using a Thevenin equivalent
(Fig. 2.7) where
) )
R3 R4
Vth = V0 − and Rth = R1 ||R3 + R2 ||R4 . (2.7)
R1 + R3 R2 + R4

Using the voltage divider equation, the voltage across the meter, the output voltage,
is
) )
R5
Vout = Vth . (2.8)
Rth + R5

With some effort, (2.8) can be written in terms of the original resistors in the form

R2 R3 R5 − R1 R4 R5
Vout = V0
R1 R2 R3 + R1 R2 R4 + R1 R2 R5 + R1 R3 R4 + R1 R4 R5 + R2 R3 R4 + R2 R3 R5 + R3 R4 R5
, (2.9)

where the denominator includes all the possible combinations of three resistors except
R1 R3 R5 and R2 R4 R5 .

3 The circuit was originally described by Samuel H. Christie. Charles Wheatstone, a British Scientist
and inventor, popularized it, giving proper citation to Christie, however Wheatstone’s name became
attached to the circuit.
2.2 The Wheatstone Bridge 35

Fig. 2.6 The Wheatstone bridge circuit as it is often drawn is shown in (a), where the device “M”
is a meter. The circuit as it might be drawn as the first step in an algebraic solution is shown in (b).
The meter is replaced with an equivalent resistance, R5

Fig. 2.7 The bridge circuit


when all but R5 (the meter) is
replaced with its Thevenin
equivalent

The Wheatstone bridge is often used for a null measurement where the bridge is
balanced so that V out = 0. To do that, at least one of the resistors should be variable
in a known way. The balance occurs when R2 R3 – R1 R4 = 0, or equivalently,

R1 R2
= . (2.10)
R3 R4

If three of the resistance values are known, then it is clear that the fourth can then be
determined. Experimentally two of the values do not need to be known. Suppose R1
and R2 are fixed values, R3 is adjustable and R4 is the unknown to be determined. The
bridge is first balanced by adjusting R3 until no current flows through the meter. Then
R4 = R2 R3 /R1 . Now the fixed resistors R1 and R2 are swapped and R3 is once again
adjusted to rebalance the bridge. Call the new value of R3 , R3' . Then R4 = R1 R3' /R2 .
Taking the product of these two results gives R42 = R3 R3' . Values for R1 and R2 are
not required to find R4 and any uncertainty in those values will not contribute to the
uncertainty of the result—quite the experimental convenience.
By taking appropriate derivatives (e.g., such as dV out /dR3 ) evaluated at the balance
point, it can be shown that near the balance condition, V out is most sensitive to a
change in the value of any of the bridge resistors when R1 ≈ R2 ≈ R3 ≈ R4 .
Null measurements can be made to be very sensitive since it is much easier to
sense the difference between zero and non-zero values than it is to sense the difference
between two non-zero, but similar, values.
36 2 Additional Theorems

Fig. 2.8 A stick figure as a


simplified representation of
the original Wheatstone
bridge circuit of Fig. 2.6a

2.2.1 Wheatstone Bridge “Hieroglyphics”

Sometimes scientists come up with alternate descriptions and/or alternate ways of


describing problems in order to help understand them better [4].4 Here we show how
one might do this for the Wheatstone bridge. What follows is by no means a standard
notation. Here the bridge solution is shown in a different way to see what is special
about the terms that are not included.
Representing the original bridge circuit with the stick figure shown in Fig. 2.8,
each product term is described by drawing the corresponding line segments. Hence,
the term R1 R3 R4 would be represented by the symbol , R2 R3 R5 by the symbol ,
and so on. Then the output voltage (2.9), is given by the picture equation

(2.11)

where the denominator, as before, includes of all the combinations of three resistors
except two. The two missing combinations correspond to the symbols and . Can
you see what is different about these missing “terms” compared to the others?5
This pictorial way of representing the equation also makes it easy to remember,
should that ever be a concern—there is the difference between the “lightning bolts”
in the numerator and all combinations of three except the “satellite dishes” in the
denominator.

2.3 The Reciprocity Theorem

The reciprocity theorem is a general result from electricity and magnetism. Only a
limited statement of that result, which can be applied to circuit analysis, is presented
here.

4 For example, the so-called “Feynman diagrams” are often used in advanced quantum mechanics.
They are a well-established way of representing complex and abstract mathematical expressions
using pictures.
5 If you want to learn more about this, a course on graph theory might be of interest. In fact, a

method to solve arbitrary arrays of equal valued resistors using graph theory is presented by [3].
2.3 The Reciprocity Theorem 37

Fig. 2.9 The reciprocity theorem says that the current through the ammeter will be the same for
these two circuits

In a single source circuit, a voltage source can be considered to be providing an


“excitation.” Somewhere else in the circuit, there is a current, a “response,” to that
excitation. With a linear circuit, the size of the response is always proportional to the
size of the excitation.
The reciprocity theorem says that:
In a (linear) single voltage source circuit, the ratio of the source voltage to a current measured
somewhere else in the circuit, is the same when the positions are interchanged.

Consider a (possibly hypothetical) box with any number of passive linear circuit
elements (passive here meaning there are no sources) inside. The elements are
connected in any way and there are four wires sticking out of the box. The reci-
procity theorem says that if an ideal voltage source is connected to one pair of wires,
and the current measured with an ideal ammeter connected across the other pair,
you will get the same reading on the ammeter if you switch the two devices. This is
illustrated in Fig. 2.9.
As a concrete example, the two circuits in Fig. 2.10 have the same 6 V “excitation,”
and so will have the same reading on the ammeter (recall that an “ideal ammeter”
measures the current but otherwise looks like a wire). That is, reciprocity gives
I2 = I1' . Hence, I 2 can be found by finding I1' instead. The only trick here is to keep
track of minus signs. Note, however, that the theorem does not say that I1 = I2' .6
The reciprocity theorem may allow the solution of one problem by the substitution
of another (hopefully simpler) problem. The reciprocity theorem also works with
current sources and ideal voltage measurements swapped in the same way.

2.3.1 Example—R-2R Ladder with Sources

Consider the “ladder” circuit with four voltages sources shown in Fig. 2.11. The goal
here is to determine V out . Note that if the current I (shown) can be determined, then
V out = V 4 − (2R)I.

6In the example circuits shown, the reading on the ammeter is 0.55 mA. The current out of the
battery is 1.27 mA for the circuit on the left, and 1.09 mA for the circuit on the right.
38 2 Additional Theorems

Fig. 2.10 A simple example of the reciprocity theorem. The current through the ammeter will be
the same for these two circuits. Note that the current supplied by the battery is not the same

Fig. 2.11 A type of ladder circuit with multiple sources

With considerable effort, this circuit can be solved using circuit reduction or by
solving equations generated using Kirchhoff’s laws. The solution below is not only
simpler, but also easily generalizes to the case of a much larger ladder circuit with
N sources. The solution relies on superposition, reciprocity, and parallel and series
resistor equivalents.
First consider a calculation with the source V 4 only, with all of the other sources
“turned off.” Since they are all voltage sources that means they are replaced with
wires (Fig. 2.12). From the symmetry, it is clear that the current I 2 splits into two
paths with I 1 = I 2 /2.
The two resistors on the left of Fig. 2.12 are in parallel and hence can be replaced
with an equivalent, 2R||2R = R, shown in Fig. 2.13a. Once that is done, the two
resistors now on the left are in series giving an equivalent resistance of 2R. Hence,
I2' = I2 = I3 /2. This process is repeated until there is only one loop left as shown
in Fig. 2.13b. So I 4 = V 4 /(4R), and going backward, the currents everywhere from
the single source V 4 are I3 = I3' = I4 /2 = V4 /(8R), I 2 = I 3 /2 = V 4 /(16R), and I 1
= I 2 /2 = V 4 /(32R).

Fig. 2.12 The ladder circuit with only the source V 4 , along with reciprocity and superposition, can
be used to determine V out for the original circuit of Fig. 2.11
2.3 The Reciprocity Theorem 39

Fig. 2.13 Circuit reduction


is used to find the current
from the one source V 4 of
Fig. 2.12

Reciprocity says that the location of the voltage source and the location where the
current is measured can be switched and the resulting value (at that location only)
will be the same. Hence, suppose the source V 4 is moved to another location, such
as the position for source 2, as shown in Fig. 2.14. Then the current on the right will
be the corresponding current computed above. For the position shown it will equal
I 2 = V 4 /(16R). From this result it can be concluded that with the original source V 2
alone, there will be a current on the right equal to V 2 /(16R) in the direction shown.
By solving for the source V 4 alone, the solution for V 2 alone is also obtained. To
see how powerful the theorems are, try computing that result using one of the other
techniques!
Using the same process for each of the original sources, the total current on the
right of the original circuit can be computed using superposition and the solution
above for the source V 4 alone. Note that there is a sign difference between the current
contributed to I from V 4 and those arising from the other source locations. Adding
the currents from each source considered alone, the resulting total current is

I = (V4 /4 − V3 /8 − V2 /16 − V1 /32)/R, (2.12)

and hence

Vout = V4 − 2I R = V4 /2 + V3 /4 + V2 /8 + V1 /16. (2.13)

Using induction, one can conclude that a similar circuit with N sources will have
an output

1 E Vk 1 E
N N
Vout = = Vk 2k . (2.14)
2 k=1 2 N −k 2 N +1 k=1

Circuits similar to this example can be useful for digital to analog (D to A)


conversions, considered later. In that case, each of the voltage sources takes on one

Fig. 2.14 Reciprocity says


that if the source is moved,
the corresponding current
also moves, as shown
40 2 Additional Theorems

of two values: for example, 0 V representing a binary digit 0 and a fixed non-zero
value, such as 3.3 V or 5 V, representing a binary digit 1.

2.4 Delta-Y Conversion

The Delta-Y conversion (sometimes called Delta-Wye, Delta-T, or pi-T) provides


another tool that can be useful for circuit reduction. It can also be useful if, for
some application, the necessary component values for one configuration become
impractical (which can often happen for certain filter and attenuator circuits). The
values for the other configuration may be more reasonable and will produce the same
result—they are equivalent as far as the rest of the circuit is concerned.
Referring to Fig. 2.15, the combination on the left is the “delta configuration”
(as often drawn, it is an upside-down /\). The combination on the right is the “Y-
configuration”. One can be replaced with the other using the Delta-Y conversion.
That is,

R1 R2 + R1 R3 + R2 R3 RA RB
RA = R1 =
R3 R A + R B + RC
R1 R2 + R1 R3 + R2 R3 R A RC
RB = R2 =
R2 R A + R B + RC
R1 R2 + R1 R3 + R2 R3 R B RC
RC = R3 = (2.15)
R1 R A + R B + RC

This transformation should not be memorized, but be aware that it exists and refer to
it when needed. The biggest challenge when using this transform is bookkeeping—
carefully use labels on the resistors and the connection points, as shown, to be sure
everything is correct.
Example Equivalent of Bridge Resistors
As an example, find the current I for the bridge circuit of Fig. 2.16a. Note that there
are no parallel or series resistors. Nevertheless, the Delta-Y conversion can be used
for circuit reduction.

Fig. 2.15 The delta (left)


and “Y” (right)
configurations for three
resistors. The delta-Y
transformation can be used
to replace one with the other
2.4 Delta-Y Conversion 41

Fig. 2.16 As an example, the delta-Y transformation is used to reduce the bridge circuit in (a) to
a circuit easily solved using parallel and series resistor equivalents. The steps for the transform are
shown in (b) and (c)

First, redraw and carefully label the circuit. There are several ways to do this. One
possibility is shown in Fig. 2.16b. Now transform the Delta (RA , RB , and RC ) to a
Y, as shown in Fig. 2.16c. From there, it is straightforward to use series and parallel
resistor equivalents to find

Req = (2k + 200)||(1k + 600) + 600 = 1526u. (2.16)

The current is then I = 6 V/1526 u = 3.93 mA. Try this by replacing the top
resistors instead of the bottom resistors. The result should be the same.
Can you now convince yourself that a box containing any number of resistors,
connected in any way, and with three leads coming out of it, such as illustrated in
Fig. 2.17, can always be replaced by a delta (or Y) circuit if appropriate values are
found? If so, how might you go about finding those values? If not, can you find an
example where it cannot be done?

Fig. 2.17 Is it always possible to convert an arbitrary box with three external leads (left) to a simple
delta (right)?
42 2 Additional Theorems

2.5 The Kelvin Bridge

The Kelvin bridge is used to measure a very small resistance, where the resistance of
the connections and wires might be large compared to the resistance to be measured.7
This occurs, for example, when the resistance of a short wire or a piece of metal is
to be measured. Of importance for these determinations is the use of a four-wire
resistance measurement. The circuit analysis strategy here is to use circuit reduction,
in particular the /\-Y conversion, to transform the Kelvin bridge to a Wheatstone
bridge where the solution is already known (see Eq. (2.9)).
A schematic for the Kelvin bridge is shown in Fig. 2.18, where
• R1 = the sample to be measured
• R2 = a small valued standard resistor
• R7 = resistance of a wire in series with contact resistances at D and E (all unknown,
but possibly of comparable size or even much larger than R1 )
• Rg = meter or detector
• R3 , R4 , R5 , R6 , and R8 = larger valued resistors.
Four separate contacts are made to R1 and R2 , such as is illustrated in Fig. 2.19, and
there will be some small, but unknown resistance at each of those contacts. Aside
from those included in R7 , the contact resistance at each connection point is small
compared to the series fixed resistance attached to it (e.g., the resistance at contacts
C and F are small compared to R8 , at G it is small compared to R3 , etc.).
Now use the /\-Y transformation to convert the “/\” formed by R4 , R6 and R7 to
a “Y”. Defining a value D, D = R4 + R6 + R7 , for convenience, the transformed
values are

R A = R4 R6 /D
R B = R6 R7 /D (2.17)
RC = R4 R7 /D.

The circuit is then converted as shown in Fig. 2.20, where RD = R1 + RC , RE = Rg


+ RA , RF = R2 + RB , and the current through RE is the same as the current through
Rg .
The null condition can then be found as for the Wheatstone bridge (2.10).
Rewriting the equation for the null condition for the equivalent Wheatstone bridge
in terms of the original resisters of the Kelvin bridge, the null condition is

7 The Kelvin bridge is named for the physicist known as Lord Kelvin, also known as William
Thomson (1824–1907). Sometimes you will see this same bridge referred to as the Thomson bridge.
Thomson was knighted by Queen Victoria for his various contributions; most notably work on the
transatlantic telegraph. Sir William Thomson adopted the title Baron Kelvin (of Largs) partly for
political reasons, using the name of the Kelvin river which flowed near his laboratory. He was
elevated to the House of Lords where people then began referring to him as Lord Kelvin. At least
that’s the way the story goes.
2.5 The Kelvin Bridge 43

Fig. 2.18 The Kelvin bridge circuit, which can be used to determine smaller resistance values
precisely

Fig. 2.19 An illustration showing how the four separate contacts might be made (a) to a standard
resistor and (b) a wire to be measured. The letters correspond to those in Fig. 2.18

Fig. 2.20 To solve the Kelvin bridge, the delta-Y transformation is used to convert it to a Wheatstone
bridge, where the solution is already known

(R4 + R6 + R7 )(R2 R3 − R1 R5 ) + R7 (R3 R6 − R4 R5 ) = 0. (2.18)

The null can be found without knowledge of the unknown small value R7 by
simultaneously satisfying both (R3 R6 – R4 R5 ) = 0 and (R2 R3 – R1 R5 ) = 0. These
two conditions are simultaneously satisfied when

R1 R3 R4
= = . (2.19)
R2 R5 R6
44 2 Additional Theorems

Note that such a four-lead resistance measurement technique is used for all
measurements of low value resistances, whether or not the Kelvin bridge is used.
Two of the connections supply the current through the sample and the other two to
measure a voltage difference. When using comparative words such as “low,” it is
always important to ask, “low compared to what?” Here “low value” means compa-
rable to or smaller than the resistance of the wires and electrical contacts. For some
measurements, such as to determine soil conductivity, a low value may not be partic-
ularly small in absolute terms—it is not as easy to connect wires to soil as it is to
connect wires to other wires.

2.5.1 Additional Application—Resistivity of Lamellae

A thin sheet, foil, or layer of material with uniform thickness, a “lamella,” can be
connected to the Kelvin bridge using four contacts such as illustrated in Fig. 2.21.
The current introduced through two of the contacts will create a voltage that can be
measured between the other two. The ratio of the measured voltage to the supplied
current will yield a value with units of ohms, and can be treated as a resistance. Of
course, the value for that resistance will depend on where the contacts are made. L.
J. van der Pauw showed that two such measurements could be combined to find the
resistivity of the material.8 The resistivity is a material property, like density, which
is independent of size and shape.
In order to use the van der Pauw technique, the lamella should have uniform
electrical properties, be of uniform thickness, and the sample should be “simply
connected,” meaning there are no holes. A resistance measurement is made with the
current connections (C and D) and the voltage connections (G and H) on the edge of
the sample and adjacent to each other (see the labels in Fig. 2.21). Then all connections
to the Kelvin bridge are rotated by one place and another measurement is made (see
the labels in parentheses in Fig. 2.21). Call the result of the first measurement RA ,
and the second RB . Then the resistivity, ρ, is given by

π d RA + RB
ρ= f (r ), (2.20)
ln 2 2
where f(r) is the solution to the wonderfully complicated equation
| |
r −1 f (r ) exp(ln 2/ f (r ))
= cosh−1 , (2.21)
r +1 ln 2 2

for which no analytic solution is known. Using r = RA /RB or RB /RA , whichever is


larger, only values for r ≥ 1 need to be known. It is not difficult to make a plot of f (r),

8This was done using conformal mappings in the complex plane—a topic from Complex Analysis.
See van der Pauw (1958/59).
2.5 The Kelvin Bridge 45

Fig. 2.21 To determine the resistivity of lamella, four contacts are used around the edges. The
letters correspond to the connection points in the Kelvin bridge (Fig. 2.18). A second measurement
is made using the connection points identified in parentheses

Fig. 2.22 A graphical


representation of van der
Pauw’s function, f (r)

such as Fig. 2.22, or to create a table of values. A number of approximate relations


have also been developed that can be used to find f (r) to good accuracy.
Notice that the actual shape of the sample is not important for the van der Pauw
measurement as long as the connections are made at the edge and the sample is simply
connected. A symmetric sample with symmetric connections, such as a square foil
with contacts at the corners, will have RA = RB and hence r = 1. Noting that f (1) =
1, the measured resistance will be

2ρ ln 2
RA = RB = , (2.22)
πd
which depends on the thickness and resistivity of the foil, but is independent of the
size of the square. Such a result can thus be expressed using units such as “ohms per
square.”9 For layered materials of known quality, a measurement of ohms per square

9“Ohms per square” usually refers specifically to measurements of the resistance from one side of a
square sheet to the opposite side, ignoring any contact resistance, and is also referred to as the sheet
46 2 Additional Theorems

can be used to determine the thickness of the layer. For a material of known thickness,
such a measurement can be used to characterize the quality of the material. The latter
measurement can be used, for example, to characterize the quality of semiconductor
wafers used during the production of integrated circuits.

2.6 Problems

• Find the Thevenin equivalent for the circuit of Fig. 2.P1, including all of the
components except the 2 k resistor on the right.

Fig. 2.P1 Problem 1

• (a) What is the Thevenin equivalent for the piece of a circuit shown in Fig. 2.P2a?
(b) What values should be chosen for resistors R1 , R2 , and R3 so that the piece of
a circuit shown in Fig. 2.P2b is equivalent to the circuit of part a?

Fig. 2.P2 Problem 2

• A less than ideal voltage source is connected to a voltmeter (which can be taken to
be ideal) and the output voltage is measured to be 9.0 V. A variable resistor is then
placed across the line, in parallel with the voltmeter. When the resistor is adjusted
to 600 u the reading on the voltmeter is exactly half of the first measurement. If the
voltage source is to be modeled using a Thevenin equivalent, what are appropriate
values for V th and Rth ?

resistance. That value can be derived from the Van der Pauw method used at the corners by dividing
that result by a factor of 2 ln2/π. For more information about 4-probe resistance measurements
using alternate geometries, see [2].
2.6 Problems 47

• (a) What is the equivalent resistance for the infinite array of equal valued resistors
shown in Fig. 2.P4a? (b) Compare the result from (a) to that of the finite array of
resistors in Fig. 2.P4b.

Fig. 2.P4 Problem 4

• Demonstrate reciprocity with a current source and voltage measurement by


computing the voltage, V, for the two circuits in Fig. 2.P5.

Fig. 2.P5 Problem 5

• A clever student decided to use reciprocity to find the current I 5 in the bridge
circuit shown in Fig. 2.P6a. First, the location of the source and the location
where the current is measured are swapped, as in Fig. 2.P6b. Then the circuit is
redrawn as shown in Fig. 2.P6c. The student correctly solves for the current I =
(72/85) mA and concludes that must also equal the current I 5 from the original
bridge circuit. However, that answer is exactly a factor of 12 too large. What did
the student do incorrectly and where does the factor of 12 come from?

Fig. 2.P6 Problem 6


48 2 Additional Theorems

• What is the Thevenin equivalent for the circuit of Fig. 2.P7?

Fig. 2.P7 Problem 7

• (Challenge Problem) Use symmetry and circuit theorems to determine the effec-
tive resistance for the following: (a) The circuit of Fig. 2.P8a, a square array
with diagonal connections, measured across adjacent corners, (b) the circuit of
Fig. 2.P8b, an infinite array of the squares from A. (Hint: if you end up using
reams of paper for either of these problems, you are not thinking like a scientist).

Fig. 2.P8 (a) Problem 8a, (b) Problem 8b


References 49

References

1. D. Atkinson, F.J. Van Steenwijk, Infinite resistive lattices. Am. J. Phys. 67, 486–492 (1999)
2. I. Miccoli, et al., The 100th anniversary of the four-point probe technique: the role of probe
geometries in isotropic and anisotropic systems. J. Phys. Condens. Matter 27, 223201 (2015)
3. L.W. Shapiro, An electrical lemma. Math. Mag. 60(1), 36–38 (1987)
4. L.J. van der Pauw, Philips Research Reports, vol. 13, No. 1, pp. 1–9. Philips Tech. Rev. 20(8),
220–224 (1958/59)
Chapter 3
Complex Impedances

With the basic theorems in place for the simplest devices—ideal power sources,
resistors and wires—some additional linear devices are now considered. In particular,
capacitors, inductors, and time-dependent sources are introduced. In some cases, it
will prove most convenient to describe voltage-current relationships using imaginary
numbers. First, what constitutes a linear device is defined. Then it is shown how
imaginary values might arise, why such values may be useful, and how they are
used. Since, by definition, nothing measured is ever imaginary, it is important to
understand how to interpret these imaginary values for real world results, as well as
to know when they will not work.

3.1 What is a Linear Device?

A device is considered linear if its current–voltage characteristic is linear in the same


way that linearity is defined for mathematical functions (and operators). This does
not necessarily mean that the current–voltage characteristic can be plotted as a simple
straight line.
Assume the current through a (two-lead) device can be described as a function
of the voltage across the device. That is, it is the case that I = I(V ). If the device is
linear then the function I(V ) always obeys

I(aV1 + bV2 ) = aI(V1 ) + bI(V2 ), (3.1)

where a and b are any (unitless) real numbers and V 1 and V 2 are any two voltages.
For example, it is clearly true that a resistor is linear. For a resistor, I(V ) = V /R,
and
1 V1 V2
I(aV1 + bV2 ) = (aV1 + bV2 ) = a +b = aI(V1 ) + bI(V2 ). (3.2)
R R R

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 51


B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_3
52 3 Complex Impedances

Some other devices are also linear. Recall from electricity and magnetism that for
a capacitor, C, with charge Q,

dQ d(C V ) dV
I(V ) = = =C , (3.3)
dt dt dt
so
d d V1 d V2
I(aV1 + bV2 ) = C (aV1 + bV2 ) = aC + bC = a I(V1 ) + b I(V2 ),
dt dt dt
(3.4)

and therefore, the capacitor is a linear device. Similarly, an inductor is a linear device.
Suppose, however, there were a device where

I(V ) = D exp(αV ), (3.5)

where D and α are constants with appropriate units. Then

I(aV1 + bV2 ) = D exp(α(aV1 + bV2 )) = D exp(αaV1 ) exp(αbV2 )


= D(exp(αV1 ))a (exp(αV2 ))b
= (I(V1 ))a (I(V2 ))b /D /= aI(V1 ) + bI(V2 ), (3.6)

and this is not a linear device. Note that if I(V ) is a linear function, then V(I) will
also be a linear function (if the function exists) and vice versa.

3.2 Some Vocabulary

For success in any endeavor, it is most important to know what you are talking about.
Thus, before proceeding, it is useful to take a little space to define a number of
terms and abbreviations that will occur throughout electronics and the rest of this
text. The definitions here represent common use of these terms within the context of
electronics. These are:
• DC: An abbreviation for “direct current,” also written as D.C., d.c., dc, etc. DC
is used to indicate voltages and currents that can be considered constant in time.
• VDC: An abbreviation for “Volts DC,” that is, a constant voltage.
• AC: An abbreviation for “alternating current,” also written A.C., a.c., ac, etc. AC
is used to indicate voltages and currents that vary in time. Often it is implied
that the time variation is periodic, with a time average of zero, and also that it is
sinusoidal.
3.3 Passive Linear Circuit Elements with Two Leads 53

• VAC: An abbreviation for “Volts AC,” that is, a time dependent (usually sinusoidal)
voltage.
• Frequency, f : Used for periodic signals, the frequency, in hertz (Hz), is the
number of signal repeats, that is “cycles,” per second. In older literature, the units
may be written “cycles per second” or cps, rather than hertz.
• Angular frequency, ω: Principally used for sinusoidal periodic signals, angular
frequency is the frequency expressed in radians per second, where there are 2π
radians for each full repeat cycle. Hence, ω = 2πf .
• Period: The repeat time for a periodic signal, often symbolized using “T.” Related
to the frequency in hertz, f , by T = 1/f, and the angular frequency, ω, by T =
2π/ω.
• RMS: An abbreviation for “Root Mean Square.” Also written “rms.” The root
mean square is the square root of the average of a series of values squared (or
a function squared). This is a useful way to characterize AC signals since, for
example, power depends on the voltage or current squared. For a sinusoidal
signal, the RMS voltage is the amplitude of the sine wave divided by the square
root of 2.
• VRMS: An abbreviation for “Volts RMS,” also written as VRMS, Vrms, etc. This
is the usual unit of volts (V) for a situation where the RMS value is being reported.
That is, 115 VRMS would mean that one has an AC signal where the time average
gives a root mean square value of 115 V. It is used (sometimes) to make sure that
the reader knows the value is an RMS value. In electronics, “RMS” is often left
off, but is implied, when AC values are reported. For example, 115 VAC usually
implies that 115 is the RMS value.
• Passive Device: A device that does not require a separate power supply for normal
operation.1

3.3 Passive Linear Circuit Elements with Two Leads

The simple ideal passive linear devices in electronics are the resistor, capacitor and
inductor. Real devices are usually modeled using one or more of these three. The
rules (i.e., the current–voltage characteristics) for these devices are summarized in
Table 3.1. For all of these devices, the sign convention is such that a positive voltage
corresponds to a decrease in the electric potential (or EMF) in the direction of positive
current, illustrated in Fig. 3.1. That is, positive values are defined such that there is
a “voltage drop” in the direction of the current. The voltage-current relationships
in Table 3.1 are for ideal devices and are approximations for the behavior of real
devices.

1There are several “definitions” of passive vs. active devices. None are particularly rigorous and
none actually work in all cases. However, all agree that the resistor, capacitor, and inductor are
passive.
54 3 Complex Impedances

Table 3.1 Passive linear


Schematic Name V(I) I(V )
devices
symbol
Resistor V = IR I = V /R
R is a constant and will be in ohms
(y)
Capacitor V = I = C ddtV
{t
V0 + 1
C 0 I (t)dt
C is a constant and will be in farads
(F)
Inductor V = L ddtI I = {t
I0 + 1
L 0 V (t)dt
L is a constant and will be in henries
(H)

Fig. 3.1 For two-lead


passive devices, a positive
current is taken to be in the
direction from higher to
lower voltage

3.4 Idealized Sources

Idealized power sources are linear devices with a very simple relationship between
voltage and current. The ideal battery is one such device and was already presented.
Real power sources are often modeled using one or more ideal sources combined
with one or more passive linear devices. A summary of some ideal sources is in
Table 3.2.
There is considerable variation for the specific symbols used for sources. For
example, one might see the symbol shown for “DC Voltage Source” also used for
AC voltage sources. The reader is often expected to be able to determine what is
meant from the context.

3.5 RC and L/R Time Constants

(Note: This first section should be a review of material from introductory electricity
and magnetism).
For both the capacitor and inductor, something interesting will happen only if the
voltage and/or current is changing in time. If neither change with time, the derivatives
3.5 RC and L/R Time Constants 55

Table 3.2 A variety of ideal linear sources


Schematic Name V(I) or I(V )
Battery V = V0
A constant value, no matter what current

DC voltage V = V 0
source A constant value, no matter what current

AC voltage V = V 0 cos(ωt + ϕ)
source Where the voltage specified, V, is often the rms voltage, Vrms ,

V0 = Vr ms 2
DC current I = I 0
source A constant value, no matter what voltage

AC current I = I 0 cos(ωt + ϕ)
source Where the current specified, I, is often the rms current, I rms,

Io = Ir ms 2

(see Table 3.1) are zero and the current–voltage relationship is trivial. In this chapter,
two simple cases involving a time dependence are considered: first, a time dependence
due to a switch being opened or closed, and then the time dependence in the presence
of sinusoidal signals. The more general case can be quite complicated and will be
considered briefly in Chap. 5.
Consider the two circuits in Fig. 3.2 where there is a switch that will close (make
contact) at a certain time, t 0 . Assume the switch closes instantaneously. Before the
switch is closed, it must be that I = 0 since the circuit is not complete—it is an open
circuit. Also assume that the switch has been open a “long time” so that the voltage
across the capacitor and the inductor are both zero. Long is a relative word and what
constitutes a long time is an important question to ask, and the answer to which, for
this case, should be evident later. The more immediate question to be answered is
“what happens after the switch is closed?”
After the switch is closed, apply Kirchhoff’s voltage law (KVL), along with the
device rules, around the loops to get

V − I R − VC = 0; V − I R − VL = 0
(3.7)
−R ddtI − C1 I = 0; V − I R − L ddtI = 0,

Fig. 3.2 Simple circuits to


illustrate a RC and b L/R
time constants
56 3 Complex Impedances

where, for the capacitor circuit, a derivative was taken on both sides of the equation.
Both of these equations can be written in the form

dI 1
= − (I − I∞ ). (3.8)
dt τ
The constant, τ (a lower-case Greek tau), will have units of seconds and is known
as the “time constant.” The constant I ∞ is the current after a very long time—a “long
time” after the switch was closed things stop changing and hence the derivatives,
that is, the rates of change, become zero. It should be evident that for the capacitor
circuit shown, τ = RC and I ∞ = 0, and for the inductor circuit, τ = L/R and I ∞ =
V /R.
A general solution to the differential Eq. (3.8) is given by

I (t) = I1 e−(t−t9 )/τ + I2 , (3.9)

where I 1 and I 2 are constants. The constant t 0 is the starting time, which is often
taken to define t = 0. In fact, the solutions to all “time constant problems” will look
just like (3.9). It is possible to skip directly to this step for virtually all future time
constant problems. To determine the values of the constants, make the solution match
the problem at certain times, such as at t = t 0 and t → ∞ (though other times can
be used).
Inductors will tend to keep the current through them constant. Hence for the
inductor circuit above the initial current must be zero since it was zero before the
switch was closed. To get the solution to do that, it must be that I 2 = − I 1 . After a
long time, an inductor simply looks like a wire so I 2 = I ∞ = V /R. So, for the inductor
circuit at a time t after the switch is closed, I(t) = V (1 – e−t/τ )/R.
Capacitors will tend to keep the voltage across them constant. Hence the voltage
across the capacitor in the corresponding circuit above will start off at zero, which
is what it was before the switch was closed. Initially the capacitor looks like a wire
(or more generally a battery, and in this case a 0 V battery), and so the initial current
is (I 1 + I 2 ) = V /R. After a long time, the current through the capacitor is zero (the
capacitor becomes “charged up” and no further charge is entering or leaving). Hence,
I 2 = 0 and one gets that for the capacitor circuit shown with the switch closed at t =
0: I(t) = V (e−t/τ )/R.
For all of these time constant problems, look at the initial state of the system and
the final (long time) state of the system. Figure out the appropriate time constant
(or effective time constant) and then use the general solution above to generate a
solution that changes exponentially between the initial and final values. Examining
these two examples, it should be clear that the time constant, τ, sets the time scale.
What constitutes a long and a short time corresponds to those times that are long and
short compared to τ.
3.5 RC and L/R Time Constants 57

3.5.1 RC Time Constant Example

In the circuit shown in Fig. 3.3, the switch has been open a long time and is then
closed at t = 0. What is the voltage across the capacitor and the current, I, in the
circuit after the switch is closed?
Initial Values First, look at the circuit just before the switch is closed. The battery
is disconnected and there is no source of power. The capacitor will have discharged
(through the 3 k resistor) and so the voltages across each of the two resistors, and
across the capacitor, will be zero. Just after the switch is closed, the capacitor will
try to keep the voltage across it constant, so the initial voltage across the capacitor,
and hence also across the 3 k resistor, will stay zero (its starting value) immediately
after the switch is closed. So at t = 0 the entire 9 V is across the 2 k resistor, and the
initial current is I = 9 V/2 k = 4.5 mA.
Final Values Now look at the circuit a long time after the switch is closed. After a
long time, the capacitor will be fully charged to its “final” value, the current through
the capacitor will be zero, and the current travels through the two resistors in series.
The current after a long time is I = 9 V/(2 + 3)k = 1.8 mA. This means that the
voltage across the 3 k resistor is 1.8 mA × 3 ky = 5.4 V, and that is, of course, equal
to the (final) voltage across the capacitor.
Time Constant For this problem, the time constant when the switch is closed
is required. Any fixed power sources will have no effect on an RC time constant.
Fixed voltage sources can be replaced with wires (a “0 V source”) and fixed current
sources with open circuits (a “0 current source”—see “superposition” in Chap. 1)
and this will not change the result for this part of the calculation. Now ask, what is
the equivalent resistance as seen by the capacitor? In this case the capacitor “sees”
the two resistors in parallel giving an effective resistance of (6/5) ky, so the RC time
constant is τ = (6/5) ky × 10 μF = 12 ms. Make sure to keep track of the powers
of ten carefully and correctly.
Now put it all together The current for this problem will look like I (t) =
I1 e−(t−t0 )/t + I2 with τ = 12 ms, t 0 = 0, I(0) = 4.5 mA and I(t → ∞) = 1.8 mA.
So it must be that I 2 = 1.8 mA and I 1 = (4.5 – 1.8) mA = 2.7 mA.
For time constant problems, the voltage is found using a similar approach. The
voltage across the capacitor will look like V C (t) = V 1 e–t/τ + V 2 , and for this problem
V C (0) = 0 and V C (t → ∞) = 5.4 V. Hence, V 2 = 5.4 V and so V 1 = –5.4 V.
() ( () (
Solutions : I (t) = 2.7e−t/12ms + 1.8 mA; VC (t) = 5.4 1 − e−t/12ms V.

Fig. 3.3 Circuit used for an


RC time constant example
58 3 Complex Impedances

Now try to work the problem with the same circuit, but where the switch was
closed for a long time and then opened at t = 0.

3.6 Capacitors and Inductors with Sinusoidal Sources

Recall that Fourier’s theorem and superposition say that any time-dependent source
can be represented as a sum of sinusoidal sources. The sum of the individual solutions
is the solution to the sum. Knowing what happens to each single sinusoidal contribu-
tion by itself is sufficient knowledge to be able to treat all time-dependent sources,
so a treatment based solely on sinusoidal time dependence is surely warranted. To
learn more about Fourier’s theorem, see the Appendix.
Consider a circuit consisting of a capacitor and an AC voltage source (Fig. 3.4a),
where the current through the capacitor is easily computed to be.

dV
I(V ) = C = ωC V0 cos(ωt) = ωC V0 sin(ωt + 90◦ ). (3.10)
dt
The current looks like a constant (ωC) multiplied by the original voltage, but with
an additional “phase shift” of 90°. Such a phase shift is equivalent to a fixed shift in
time. In this case, the voltage seems to be behind, or “lagging,” the current—when
t = 0, the argument inside the sine function is zero for the voltage, but is already
larger than zero for the current.
Similarly, for an inductor connected directly to an AC voltage source (Fig. 3.4b),
the current is found to be
−1 1
I(V ) = cos(ωt) = sin(ωt − 90◦ ), (3.11)
ωL ωL
where the voltage is ahead of, or “leads,” the current by 90°—when the argument
inside the sine for the current is zero, it must be the case that ωt > 0, so the argument
inside the sine used for the voltage is already positive.
Hence, the amplitude of a sinusoidal voltage across (V ) and the corresponding
sinusoidal current (I) through the capacitor (C) and inductor (L), can be symbolically
described as

Fig. 3.4 The simplest circuits with a sinusoidal AC source for (a) a capacitor and (b) an inductor
3.6 Capacitors and Inductors with Sinusoidal Sources 59

1
VC = IC ∠ − 90◦
ωC
VL = ωL I L ∠90◦ , (3.12)

where the sinusoidal function is not shown, but implied, and “∠” means to shift the
sinusoid by the phase angle that follows. With this notation, Ohm’s law for resistors
might be written V R = RI R ∠0°, however since a shift by 0° is to do nothing, “∠0°”
is usually omitted when it occurs.
Phase angles are always relative to some reference; they are phase differences. A
phase angle is used to characterize any shifts in time of sinusoidal signals relative
to one another. A signal that is delayed in time (“lagging”) compared to a reference
has a negative phase angle compared to the reference. Since there are 360° (or 2π
radians) in one complete period, a phase angle of 90° (or π/2 radians) corresponds
to a shift of 1/4th of a period, or a time equal to T/4. The general case is illustrated
in Fig. 3.5.
Now the question remains as to what to do with these relationships for circuits that
are more complicated. That is, how are these results used in circuit analysis and how
are the calculations performed? To answer this, first consider the following approach
to solve a simple circuit.

Fig. 3.5 Showing the offset


in time, /\T, associated with
a phase shift /\ϕ. The two
signals have been offset in
the vertical direction for
clarity
60 3 Complex Impedances

3.7 Superposition and Complex Impedances

Complex numbers are those that involve a sum of real and imaginary numbers.
Imaginary numbers are those that involve the square root of minus one. If needed,
a review of complex numbers and complex number arithmetic is included in the
Appendix.
It will be most convenient to consider complex numbers and complex arithmetic
for AC circuits—in particular for circuits where the source(s) are sinusoidal. There
is no such thing as an imaginary voltage or an imaginary current, however, so this
approach should be greeted with some skepticism. The purpose here is to show why
this works and to illustrate the appropriate
√ interpretation for the complex numbers.
Throughout this text the convention i = −1 is used. In the literature, particularly
for electronics, sometimes “j” may be used in place of “i’.
Why and how complex values might be useful is illustrated using the following
example. Consider the circuit in Fig. 3.6a which has a sinusoidal voltage source of
frequency f = 10 Hz (so ω = 2πf = 62.8 rad/s). Using the mathematical identity,

1 () iωt (
cos(ωt) = e + e−iωt , (3.13)
2
One can, on paper, draw the circuit with two sources, each of which produces a
complex voltage (though the sum is always a real value). See Fig. 3.6b. Using the
superposition principle, the total current in the circuit can be found by considering
each of these voltage sources separately, with the other “off,” and then summing the
results, even though each source, by itself, can never actually exist.
Start by looking at the circuit of Fig. 3.7a, where three branch currents have
been labeled. Writing down equations obtained by applying Kirchhoff’s laws to this
circuit,

I1 = I2 + I3
2.5ei ωt − 3I1 − 2I2 ={ 0 (3.14)
2.5ei ωt − 3I1 − 20μF
t
t→−∞ I3 dt = 0,
1

Fig. 3.6 The circuit shown in (a) is solved using mathematical identity to split the source into two
sources, as shown in (b)
3.7 Superposition and Complex Impedances 61

Fig. 3.7 The circuit of Fig. 3.6a is solved considering one source at a time. The two solutions will
be identical except for the replacement of “i” with “ − i”

where voltages are in volts, and currents in mA. The integral yields the total charge
that has accumulated on the capacitor. Now if the sinusoidal source has been on for a
long time, all the currents will also be sinusoidal with the same frequency.2 So write
Ik (t) = Ik0 ei ωt , k = 1, 2, 3, and these equations become

1
I10 = I20 + I30 2.5 − 3I10 − 2I20 = 0 2.5 − 3I10 − I30 = 0, (3.15)
(iω)20μF

where the capacitor has been assumed to be initially (i.e., at t → − ∞) uncharged.


Note that there is no longer any time dependence in these equations. All the time
dependence has been factored out. For this example, ω = 2πf = 62.8 s−1 , so the last
equation becomes

2.5 − 3I10 + 0.8i I30 = 0. (3.16)

Solving these equations yields (in mA)

I10 = (0.73 + 0.15i ) = 0.75e+0.21i


I20 = (0.15 − 0.23i ) = 0.28e−0.99i
I30 = (0.58 + 0.39i ) = 0.69e+0.59i , (3.17)

and so the currents from this first (complex) voltage source are

I1 (t) = 0.75e+i(ωt+0.21) ; I2 (t) = 0.28e+i(ωt−0.99) ; I3 (t) = 0.69e+i(ωt+0.59) .


(3.18)

Now look at the second source by itself with the first source “off,” shown in
Fig. 3.7b, and analyze this circuit in the same way. The only difference is equivalent
to replacing i with − i everywhere it appears. This means that the result will be the
same as the previous result, but with each i replaced with − i. Thus, for this second
source

2 There is no other frequency that it could be.


62 3 Complex Impedances

I1 (t) = 0.75e−i(ωt+0.21) ; I2 (t) = 0.28e−i(ωt−0.99) ; I3 (t) = 0.69e−i(ωt+0.59) .


(3.19)

To find the solution to the original problem, add these two solutions together to
get

I1 (t) = 0.75(e+i(ωt+0.21) + e−i(ωt+0.21) ) = 1.50 cos(ωt + 0.21)


I2 (t) = 0.56 cos(ωt − 0.99)
I3 (t) = 1.38 cos(ωt + 0.59). (3.20)

Using the abbreviated notation from above, the solutions can be written (again in
mA),

I1 = 1.50∠12◦ , I2 = 0.56∠ − 57◦ , I3 = 1.38∠34◦ . (3.21)

where the phase shifts have been converted from radians to degrees.3 The phase shift
here is relative to the original voltage source. Any current that can be measured will
always have a real value, and these solutions are indeed real valued.
The second solution did not need to be found to know what the result would be.
Doing the computation for one or the other of the complex sources provides all the
information that is required. The other solution is simply found by replacing all i’s
with − i’s, but no new information is obtained. Henceforth, there is no need to waste
time considering that second solution. Notice how the amplitude and phase shift of
the final (real-valued) result is already contained in the result for a (hypothetical)
single complex source.
The solution just shown illustrates why using complex values for sinusoidal volt-
ages and currents might be useful and why the method works to find changes in
magnitude and phases, but it is not the way to solve these problems.
The procedure can be simplified for any sinusoidal source. Complex numbers can
be used to describe an impedance, Z, between any two points in a circuit. Then a gener-
alized version of Ohm’s law, V = IZ, is used where the impedance describes not only
the relative magnitudes but also any phase shifts that may occur. The impedance is no
longer a single value since it must describe two things simultaneously: a magnitude
and a phase shift. The phase angle of the impedance represents the phase difference
between a sinusoidal applied voltage and the resulting sinusoidal current through the
device.
The complex impedance for a capacitor and for an inductor are found using the
general relationships between current and voltage and assuming the time dependence
to be proportional to eiωt . Designating the (time-independent) amplitudes with a
subscript “0,” then for capacitors

3 To convert an angle in radians to an angle in degrees, multiply by (180/π) = 57.2958.


3.8 Series and Parallel Capacitors and Inductors 63
) ) ) )
d () iωt ( () iωt ( 1 1 1
C V0 e = I0 e → i ωV0 = I0 → V0 = I0 → Z C = ,
dt C i ωC iωC
(3.22)

and for inductors:


() ( d () iωt (
V0 ei ωt = L I0 e → V0 = i ωL I0 → Z L = i ωL . (3.23)
dt
Comparing these results to those above (3.12), it is clear that multiplying by “i” is
equivalent to a 90° phase shift, and dividing by “i” (which is the same as multiplying
by “–i”) is a –90° phase shift. More generally, for this electronics application using
sinusoidal sources, multiplication by eiθ is equivalent to “∠θ.” The arithmetic of
the phase shifts is the same as the arithmetic of these exponentials. Some examples
follow.
Examples


() ◦( () ◦(
(6.3∠35
() ) × (2.1∠ − 22(◦ ) ⇒ 6.3ei35 × 2.1e−i22
◦ ◦
= 6.3 × 2.1ei(35 −22 ) ⇒ 13.23∠13◦ (3.24)
i85◦
9.6∠85◦ ◦ ◦ ◦
3∠25◦
⇒ 9.6e
33i25◦
= 9.6
3
ei(85 −25 ) = 3.2ei60 ⇒ 3.2∠60◦ .

Analyzing time-dependent signals by considering the individual frequency


components (i.e., individual sinusoidal contributions to a more complicated signal)
is to analyze the signal “in the frequency domain” rather than in the “time domain.”
For capacitors and inductors, the complex impedance is also frequency dependent.
Hence, if a complicated signal is written as a sum of sinusoidal signals, then, in some
sense, each sinusoid will experience a different impedance.
The use of complex numbers for problems involving any type of sinusoidal motion
or signal is very common and is a very useful tool, whether the problem is from
electronics, electricity and magnetism, mechanics, seismology, optics or any number
of other fields. As long as there is a linear response, and hence superposition, complex
numbers might be used to advantage when describing a time-dependent problem.

3.8 Series and Parallel Capacitors and Inductors

All the basic results and theorems presented for resistors apply more generally to
complex impedances. For example, you can easily derive the following equivalences
directly from the results for parallel and series resistors:
• Inductors in series:

Z eq = Z 1 + Z 2 = i ω(L 1 + L 2 ) = i ωL eq → L eq = L 1 + L 2 (3.25)
64 3 Complex Impedances

• Inductors in parallel:
) )
1 1 1 1 1 1 1 1 1 1
= + = + = → = + (3.26)
Z eq Z1 Z2 iω L1 L2 i ωL eq L eq L1 L2

• Capacitors in series:
) )
1 1 1 1 1 1 1
Z eq = Z 1 + Z 2 = + = → = + (3.27)
iω C1 C2 i ωCeq Ceq C1 C2

• Capacitors in parallel:

1 1 1
= + = i ω(C1 + C2 ) = iωCeq → Ceq = C1 + C2 (3.28)
Z eq Z1 Z2

Of course, these same results can also be derived directly from the time domain
relationships between voltage and current.
As an example, compare the infinite array of capacitors, Fig. 3.8a, to the infinite
array of resistors solved previously. The translational symmetry is used as before
(Fig. 3.8b). Then

1 1 1 () ( () (
= + → C1 Ceq + C2 = Ceq Ceq + C1 + C2 , (3.29)
Ceq C1 C2 + Ceq

So
2
Ceq + C2 Ceq − C1 C2 = 0, (3.30)

and then, using the quadratic formula,


) / )
1
Ceq = −C2 ± C22 + 4C1 C2 , (3.31)
2

where the plus sign must be chosen to remain physical. For the special case where
C 1 = C 2 = C, this becomes

Fig. 3.8 Using complex impedances, an infinite array of capacitors can be solved in the same
manner as an infinite array of resistors
3.9 Comments About Complex Arithmetic 65
(√ )
5−1
Ceq =C . (3.32)
2

3.9 Comments About Complex Arithmetic

Complex values can be represented using real and imaginary parts and/or a magnitude
and a phase angle. That is, a complex value can be represented using two real values,
A and B, or two real values C and ϕ where
/
A + i B = Ceiϕ ; C = A2 + B 2 ; tan(ϕ) = B/ A. (3.33)

There is a direct analogy to vector notation in mechanics, where x–y coordinates


or polar coordinates might be used. When adding or subtracting complex values, it is
most useful to use real and imaginary parts (A and B). When multiplying or dividing,
it is often most useful to use magnitudes and phases (C and ϕ). Hence being able to
convert between the two notations is essential.
An arctangent is often used when determining the phase angle from real and
imaginary parts, however a calculator or computer will provide answers in only two
quadrants. The result must be checked to make sure it is in the correct quadrant. For
example, when computing the phase angles for 3–5i and –3 + 5i using the arctangent,
a calculator returns the same answer, but the angles are really 180° apart. For most
calculators, if the real part is negative you need to add 180° to (or subtract 180° from)
the result. For example, for the values above
) ) ) )
−1 Im Part −1 −5
ϕ1 = tan → tan = −59◦ ;
Re Part 3
) )
−1 5
ϕ2 = tan = −59◦ + 180◦ = 121◦ . (3.34)
−3

To check results, it helps to plot the values in the complex plane, such as in
Fig. 3.9. The phase angle is the angle “up from the real axis.” “Up” in this case
would be counterclockwise.
For electronics, phase angles are normally reported between −80° and 180° unless
there is a particular reason to extend them outside that range. If angles are outside
that range, an appropriate multiple of 360° can be added or subtracted to get them
back in that range.
66 3 Complex Impedances

Fig. 3.9 Care must be taken when converting complex numbers to a magnitude and an angle since
the arctangent function will return results in only two of the four quadrants. Representing the values
graphically helps to identify the correct quadrant

3.10 Solving Circuits Using Complex Impedances

Here it is shown, by example, how to solve circuits with sinusoidal sources. The
approach is identical to what was used with resistors, however the arithmetic becomes
more complicated due to the addition of phase angles (i.e., complex numbers).
Example 1
It is desired to find the currents in the circuit of Fig. 3.10a. For this problem, the
solution is shown using complex impedances and circuit reduction. Kirchhoff’s laws
and equations could be used instead, if that is desired. The first step for either is
to compute the impedance for the capacitor. Aside from forgetting to make the
computation altogether, a common mistake when computing the impedance is to
use the frequency, f , rather than the angular frequency, ω, in the computation. The
difference is almost a factor of ten! The impedance of the capacitor for this example
is.
1 1
= = −800i = −i0.8k. (3.35)
iωC i (2 · π · 10Hz) · 20 × 10−6

The 2 k resistor and the capacitor are in parallel, so create an equivalent impedance
) )
Z 2k Z C (2)(−0.8i ) −1.6i 2 + 0.8i
Z eq = = k= · k
Z 2k + Z C 2 − 0.8i 2 − 0.8i 2 + 0.8i

Fig. 3.10 The circuit shown in a is solved using complex impedances and circuit reduction
3.10 Solving Circuits Using Complex Impedances 67

1.28 − 3.2i
= k = (0.28 − 0.70i )k. (3.36)
4.64
As seen by the voltage source and the 3 k resistor, the circuit looks the same as
Fig. 3.10b. Now the 3 k resistor and Z eq are in series. Hence, create another equivalent
impedance for these series components,
'
Z eq = 3k + Z eq = (3.28 − 0.70i )k
/
= 3.282 + 0.702 k∠ tan−1 (−0.70/3.28) = 3.35k∠ − 12◦ . (3.37)

Hence,
) )
5V 5
I1 = = mA∠ + 12◦ = (1.45 + 0.31i )mA. (3.38)
3.35k∠ − 12◦ 3.35

Now in the original circuit, this current was split two ways, and the current divider
equation can be used to get

ZC I1 (2k)(Z C ) Z eq
I2 = I1 = = I1
2k + Z C 2k 2k + Z C 2k
) )
0.28 − 0.70i
= 1.5∠12◦ mA
2
= (1.46 + 0.31i)(0.14 − 0.35i )mA
= (0.31 − 0.47i)mA = 0.56mA∠56◦ . (3.39)

and a similar calculation is done for I 3 . Alternatively, use I 1 = I 2 + I 3 (i.e., KCL)


to get

I3 = I1 − I2
= (1.46 + 0.31i )mA − (0.31 − 0.47i )mA
= (1.15 + 0.78i)mA = 1.4mA∠34◦ . (3.40)

It is good form at this point to do some simple checks. One possible check is to
make sure the voltages across the 2 k resistor and across the capacitor are indeed
equal. That is, compute

VR = I2 (2k) = 1.1V∠ − 56◦


VC = I3 Z C = (1.4∠34◦ )(0.8∠ − 90◦ ) = 1.1V∠ − 56◦ , (3.41)

and the results do check out. Remember that when two complex values are multiplied,
the magnitudes multiply but the angles add.
68 3 Complex Impedances

Example 2
Here it is desired to find the Thevenin equivalent for all of the circuit of Fig. 3.11a
except for the 1 k resistor. All of the theorems shown for resistors will work for
an AC analysis with linear devices. Hence, follow the same procedure used before
except now with complex impedances. The first step is to compute the impedance of
the capacitor. The angular frequency is ω = 2π(60 Hz) = 377 s−1 . Then.

1 −i
ZC = = () ( y = −i2.7 ky. (3.42)
i ωC (377) 1 × 10−6

The problem is to find the Thevenin equivalent for the portion of the circuit shown
in Fig. 3.11b.
First, find the Thevenin equivalent voltage, V th , by finding the voltage between
the two leads on the right when nothing is connected. With nothing connected, there
is no current through the 3 k resistor and hence there is no voltage change across the
3 k resistor. Thus, the voltage across the capacitor equals V th .
With nothing connected, the 2 k resistor and the capacitor are effectively in series
so you can use the voltage divider equation to find the voltage across the capacitor.
That is,

ZC −i2.7k 2.7k∠ − 90◦


Vth = 6V = 6V = 6V = 4.8V∠ − 37◦ .
Z 2k + Z C 2k − i2.7k 3.36k∠ − 53◦
(3.43)

Second, find the Thevenin equivalent impedance by turning off the voltage source
(making it a wire) and then look back into the circuit from the right. See Fig. 3.11c.
Hence,

(2k)(−i2.7k)
Z th = 3k + 2k||(−i2.7k ) = 3k +
2k − i2.7k
3k(2k − i2.7k) + 2k(−i2.7k)
=
2k − i2.7k
6 − i13.7 15∠ − 66◦
= k= k = 4.5k∠ − 13◦ . (3.44)
2 − i2.7 3.36∠ − 53◦

Fig. 3.11 The procedure to find the Thevenin equivalent of the circuit shown in a is the same as
was used for resistors, except complex impedances are used
3.11 AC Power 69

The Thevenin equivalent is then constructed using these complex values for the
voltage and impedance. In each case, the phase shift is referenced to that of the
original voltage source.

3.11 AC Power

The instantaneous electrical power, P, being delivered to, or absorbed by, a device is
given by P = VI (see Chap. 1). The time-averaged power is usually of most interest
for AC devices. For example, the brightness of a light bulb, the power of an electric
motor, or the loudness of a sound from a loudspeaker, all depend on the time-averaged
power.
For a simple sinusoidal signal, the time-averaged power can be computed using
a single cycle (all other cycles are the same). Hence, if the voltage across a device
and current through the device are given by

V (t) = V0 cos(ωt + α), I (t) = I0 cos(ωt + β), (3.45)

then the time-averaged power is given by

{T {T
1 V0 I0
P= V (t)I (t)dt = cos(ωt + α) cos(ωt + β)dt
T T
0 0
{2π
V0 I0 V0 I0
= cos(ϕ + α) cos(ϕ + β)dϕ = cos(β − α), (3.46)
2π 2
0

where the period, T = 2π/ω and the substitution ϕ = ωt was used for integration.
Similarly, the time-averages of V 2 (t) and I 2 (t) are

V02 I2
V2 = , I2 = 0 . (3.47)
2 2
In cases where the time averages of V (t) and I(t) are zero, it is often useful to
characterize their time averages using the root mean square (rms) values.4 Hence,

4 In electronics, the phrase “average of a sinusoidal signal” may refer to the average of the absolute
value of the signal. This somewhat sloppy use of the word average seems to have arisen because
some early analog AC/DC meters use the interaction between two electromagnets connected in
series to provide a needle movement. That interaction is, of course, unchanged when the sign of
the current changes and so the meter reads absolute values. Due to the needle’s mechanical inertia,
the time-average value results. For a sinusoidal signal of amplitude, A, the average of the absolute
value is 2A/π, which is slightly smaller than the rms value.
70 3 Complex Impedances
/ /
V0 I0
Vr ms = V2 = √ , Ir ms = I2 = √ , (3.48)
2 2

where in the far right of each equation, the divisor 2 is only valid for a sinusoidal
time dependence.
Using these rms values, the time-averaged power for a sinusoidal signal is then

P = Vr ms Ir ms cos φ = Vr ms · Ir ms (3.49)

where φ is the difference between the phase angles. The vector dot product notation
is used as if V and I were vectors that differ by an angle φ. When it is clear that
sinusoidal signals are present, the over-bar and subscripts are often dropped but are
understood to be present.
Alternatively, the problem can be approached using ideas from superposition.
That is, write

V0 iωt I0 i ωt
V (t) = e , I (t) = e , (3.50)
2 2
where V 0 and I 0 are complex values. The appropriate calculation is:
⎡ ⎤
{T
1
P = Re⎣ V (t)I ∗ (t)dt ⎦. (3.51)
T
0

The form in brackets often appears when using complex numbers to do computa-
tions using sinusoidal functions. It can be regarded as a kind of dot product between
two complex functions.
For AC signals, the rms voltages and currents are usually specified unless it is
otherwise stated. Thus, as an example, for an AC voltage of 115 V and a corresponding
AC current of 0.75 A with a relative phase angle of 32°, the time-averaged power is
(115 V)(0.75 A)cos(32°) = 73 W. The result is quite simple, though there is a lot of
math behind where it came from. The cosine of the phase difference, in this example
cos(32°), is referred to as the “power factor” for the circuit.
The real part (“Re”) used above for the power gives the actual power in watts. If
power is treated as a complex value, the imaginary part is called the “reactive power.”
Reactive power is delivered to a device during one-half of the cycle, and is returned
from the device during the other half. Any time the values of a voltage and a current
are multiplied, the result will have units of volts times amps. In terms of basic units,
1 V · 1 A = 1 W. However, the watt is usually reserved for actual power dissipated
or delivered, which takes into account the power factor. This is why some devices
are specified using V · A instead of watts. For example, V · A is often seen for higher
power electric motors that may have the current and voltage out of phase under some
conditions.
3.12 Condenser Microphones 71

3.12 Condenser Microphones

In some older literature, capacitors are sometimes referred to as “condensers.” That


historical name is still used on occasion. For example, a “condenser microphone”
is a capacitor designed so that the capacitance depends on changes in air pressure.
The capacitor is made part of an RC circuit and the changing voltage associated
with the changing capacitance gives rise to an audio signal. Some external voltage
source is necessary to charge the capacitor—usually supplied using a battery or using
“phantom power” from an audio control board (48 V or so is typical).
A circuit with an impedance that depends on time can prove to be difficult to
solve in general. Often it is the case that the time scales involved allow a simpler
approximate solution. The condenser microphone is an example of such a problem.
The basic setup is shown in the Fig. 3.12. The electrically conducting diaphragm
responds to changes in air pressure, changing the spacing of the capacitor plates. The
RC time constant is made long enough so that the charge on the sensing plates is
(roughly) constant in time. Hence, the voltage will change with time as the capacitor
plate spacing changes. Assuming a long RC time constant, and using the simple
parallel plate model for capacitors,

Q Qd(t) Q[d0 + /\d(t)] Q/\d(t)


VC (t) = ≈ = = V0 + , (3.52)
C(t) ε0 A ε0 A ε0 A

where A is the area of the plates, d is the spacing between the plates, d 0 is the
equilibrium spacing between the plates, and ε0 = 8.854 × 10−12 F/m.
Try solving this circuit without making any approximations. Note that when the
capacitance is time-dependent you cannot use the complex impedance derived above.
An electret microphone is also a kind of condenser microphone, though the plates
include a highly insulating layer with embedded charge. That charge will stay present
for a very long time (>100 yrs). Hence, the plates are effectively permanently charged.
The signals from an electret microphone are low so a small amplifier is usually built
into the device. Electret microphones require a power supply for that amplifier,
typically a few volts, but none is needed for the capacitive sensor. Electret micro-
phones can be very inexpensive and you will find them used for cell phones, personal

Fig. 3.12 A condenser microphone includes a time-dependent capacitance in an RC circuit. Finding


the solution to this circuit is more difficult that it may first appear
72 3 Complex Impedances

computers, and lapel microphones. They tend to be somewhat noisier than the highest
quality microphones so you will not usually see them used for the highest quality
recordings.

3.13 Problems

1. For the circuit shown in Fig. 3.P1, the switch has been closed for a long time
and then is opened at t = 0. What is (a) the voltage on the capacitor just after the
switch is opened, (b) the time constant for changes in the capacitor’s voltage,
and (c) the voltage on the capacitor a long time after the switch is opened?
Finally, (d) write an expression for the voltage on the capacitor as a function of
time for t > 0.

Fig. 3.P1 Problem 1

2. For the circuit of Fig. 3.P2, the switch is initially open (and has been for a long
time, so the capacitor is uncharged), and at t = 0 the switch is suddenly closed.
After 5 ms, the voltage across the capacitor is 2.0 V. What is C? (Hint: replace
the voltage source and resistors with their Thevenin equivalent).

Fig. 3.P2 Problem 2

3. For the circuit of Fig. 3.P3, the switch is initially closed (and has been closed for
a long time) and at t = 0 the switch is suddenly opened. Obtain an expression
for the voltage across the inductor as a function of time and√plot the results.
4. Compute the following for A = 3 + 4i, B = 7 − 3i and i = −1. Put the result
in terms of a real and imaginary part and in terms of a magnitude and phase
angle.

(a) |A| (b) A/|A| (c) A − A∗ (d) A + B (e) A · B (f) A/B


3.13 Problems 73

Fig. 3.P3 Problem 3

5. What is the magnitude and phase shift of the current through the 1 k resistor for
the circuit of Fig. 3.P5?

Fig. 3.P5 Problem 5

6. Find the Thevenin equivalent for the portion of the circuit in Fig. 3.P6.

Fig. 3.P6 Problem 6

7. For the portion of a circuit in Fig. 3.P7, what is the magnitude of the Thevenin
equivalent voltage?

Fig. 3.P7 Problems 7 and 8

8. For the circuit of Fig. 3.P7, what is the magnitude and phase of the Thevenin
equivalent impedance?
74 3 Complex Impedances

9. (Challenge Problem) suppose you have a voltage that is periodic in time but
not sinusoidal. You know it can always be expressed as a sum of sinusoidal
functions due to Fourier’s theorem, but can you treat the time-averaged power
as being the sum of the power from each of the separate sinusoids? That is, if
E E
V (t) = Vi cos(ωi t + αi ), I (t) = Ii cos(ωi t + βi ),
i i

E
will the time-averaged power be given by P = 21 i Vi Ii cos(βi − αi )? Prove
or disprove.
10. (Challenge Problem) The voltage across a capacitor, V C , is related to the magni-
tude of the net charge on the plates of the capacitor Q by the relationship V C =
Q/C, where C is the capacitance. The charge is taken to be + Q on one plate and
–Q on the other, so the total net charge on the capacitor is zero. What happens
if the charges on the two plates do not balance out? In particular, would such a
charge imbalance be of concern when analyzing electronic circuits?
Chapter 4
More on Capacitors and Inductors

Real capacitors and inductors are often less ideal than real resistors. The non-ideal
behavior can lead to some inconvenience that must be dealt with when capacitors and
inductors are measured and used. In addition, this chapter presents some interesting
new applications and results using capacitors and inductors.

4.1 Real Capacitors and Inductors

Real inductors and capacitors often have a significant resistive component. At very
high frequencies, inductors can also have a capacitive component, and vice versa. In
general, the values used to model a real component may depend on the conditions
(e.g., frequency, etc.). Any time that these imperfections are an issue, a real inductor
or capacitor can be modeled using a number of ideal components. Two very simple
models for real components, that include some resistance, are the parallel and the
series models shown in Fig. 4.1.
A “quality factor,” Q, is defined at (angular) frequency ω analogous to what is
done for the driven harmonic oscillator. As long as Q is not too small, for these
models Q is given approximately as

Rp ωL 1
QL ≈ ≈ ; QC ≈ ω R p C ≈ . (4.1)
ωL Rs ω Rs C

Q is a measure of the magnitude of the reactive to resistive impedance. A larger


value of Q means the component is closer to being ideal—it is, in some sense,
of higher quality. These approximate equations and the simple parallel and series
models work well if Q >> 1. Some devices are specified using a “dissipation factor,”
which is 1/Q, often expressed in percent, or through the use of a dissipation angle,
δ, where tanδ = 1/Q.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 75


B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_4
76 4 More on Capacitors and Inductors

Fig. 4.1 Parallel and series


models used to model the
resistance for real inductors
and capacitors

For the parallel model, a more ideal component will have a very large parallel
resistance, Rp >> ωL. For the series model, on the other hand, a more ideal compo-
nent will have a very small resistance, Rs << ωL. Whether the parallel or series model
will work best depends on the specific circumstances. In extreme cases, more elab-
orate models, for example including both a parallel and a series resistance, may be
appropriate.
A general trend is that inductors and capacitors that have larger values tend to have
a lower Q. This trend is due to the compromises that may be necessary to create the
larger values and still have a device that is usable and of reasonable size. Practically
speaking, for components often found in electronics, the presence of resistance is
often more of a concern for inductors than for capacitors.

4.2 Measuring Capacitors and Inductors

Perhaps the simplest methods to determine a capacitance or inductance is to measure


a time constant. The component to be determined is placed in series or in parallel with
a known resistance and the time constant is determined. Results from the previous
chapter are used to extract the component value. If the resistance in the device is
unknown and not negligible, multiple measurements using different known resistors
may be necessary.
If the device is to be used in an AC circuit, it is perhaps best to determine the
values using an AC circuit operating near the intended operating frequency. The more
precise methods use some sort of bridge circuit where the unknown components are
compared to known values. A generalized version of the Wheatstone bridge is one
obvious choice for such measurements. Some possible configurations are illustrated
in Fig. 4.2.
4.3 Capacitive Position Sensors 77

Fig. 4.2 A generic bridge circuit to measure a complex impedance is shown in a. Specific
implementations to measure an inductor and/or a capacitor are shown in b and c

4.3 Capacitive Position Sensors

A simple position sensor can be made using a bridge circuit and variable capacitors.
While there are many ways to do this, one way is to use a capacitive sensor consisting
of three parallel plates, illustrated in Fig. 4.3a.1 The central plate moves back and
forth as shown, and the outer two are fixed in position.
These plates can be connected in a Wheatstone bridge configuration such as shown
in Fig. 4.3b. The meter “M” provides a reading. When the center plate (2a and 2b) of
the capacitor is equidistant between the outer two plates (1 and 3), the meter reads
zero due to the symmetry of the circuit.
Using results from Electricity and Magnetism (C = ε0 A/d), the fractional imbal-
ance in the capacitance when the center plate is a small distance /\x from the center
will be /\x/d, where d is the spacing between the outer plates. Since d can be quite
small, such sensors can be used to see very small changes in position. Commercial
capacitance sensors, and associated electronics, are available that have a nominal
resolution smaller than the size of an atom.
Position sensors are also useful as a major component in force bridges, pressure
sensors, or accelerometers. The center plate is attached to a spring or cantilever and is
deflected by an applied force F 1 or by a fictitious force if you subject the entire device
to an acceleration. This is the principle used for many micro-electromechanical
systems (MEMS) based acceleration detectors used in laptop computers, some game
controllers, and as sensors for air bag deployment. In some cases, provisions are
included to supply a known adjustable force F 2 to keep the inner plate centered and
thus the force F 1 is determined directly by knowing F 2 . Devices can be constructed
to apply the appropriate force, F 2 , automatically using “feedback.” Feedback will
be discussed later.

1 For more information about capacitance sensors, see [1].


78 4 More on Capacitors and Inductors

Fig. 4.3 A sensitive


capacitive position sensor
can be constructed with two
fixed outer plates and a
movable inner plate. The
imbalance as the inner plate
moves is measured using a
bridge circuit

4.4 A Simple Circuit for Measuring Inductors

Since the resistive part of an inductor is often non-negligible, any determination of the
value of an inductance will likely need to take into account that resistive component.
That is, there are two values to measure—an inductance and a resistance. For a bridge
circuit, such as the Wheatstone bridge mentioned earlier, there will need to be (at
least) two variable components to obtain a good null on the detector. As an alternative
to the Wheatstone bridge, consider the bridge circuit shown in Fig. 4.4a, where the
impedance, Z x , is an unknown inductor, including its resistive component. Here the
inductor’s resistance has been modeled using the simple parallel model.
The resistor (R) and the two matched capacitors (C) are calibrated and adjustable.
The other two impedances will not be of consequence for the balance condition,
though will be important when determining the sensitivity of the measurement. The
voltage source is sinusoidal at an angular frequency ω.
To determine Z x , the resistor and the two matched capacitors are adjusted so that
there is no current through the detector (I d = 0). When that condition is met, there
is no voltage drop across the detector. For that condition (only) the situation is as
shown in Fig. 4.4b. Two intermediate voltages, labeled V 1 and V 2 , and a ground
symbol have been added. The ground symbol at the bottom indicates the position

Fig. 4.4 A bridge circuit for measuring an inductor a and an equivalent b when the circuit has been
balanced to have no current through the detector, Z d
4.5 Switched Capacitor Methods 79

used for the 0 V reference. Starting from the upper right (which is also at 0 V due to
the match condition) the voltage drops lead to
I )
1 1 1
V1 = I2 V2 = −R I2 (V1 − V2 ) = − (I1 + I2 ) = R+ I2 , (4.2)
i ωC iωC i ωC

which can be solved to give I 1 = − (2 – iωRC)I 2 .


Now find V 1 by starting at the bottom and going up through the unknown, Z x ,

1
V1 = I1 Z x = −I2 (2 + i ω RC)Z x = I2 . (4.3)
iωC
The far right-hand side comes from the relationship found above. Using these
results
−1 1 1
Zx = =( 2 ). (4.4)
i ωC (2 + i ω RC) ω RC 2 − 2i ωC

The parallel model for the inductor yields

1
Z x = (i ωL x )||Rx = . (4.5)
1/Rx − i /(ωL x )

Comparing to the result above for the circuit under the balance condition, the
corresponding model components are

1 1
Rx = ; Lx = , (4.6)
ω2 RC 2 2ω2 C
and the quality factor is given by Q = 1/(2ωRC). That is, once the match condition
is achieved, the component values of the inductor model are easily determined from
the (presumed known) values of the bridge resistor and the two matched capacitors.

4.5 Switched Capacitor Methods

There are a number of interesting circuits developed around the idea of a “switched
capacitor.” Only the basic principle is shown here. In practice, the switching operation
is often achieved using transistor circuits and may occur at frequencies well above
1 kHz.
Consider the circuit of Fig. 4.5a. The switches S 1 and S 2 will be alternately opened
and closed, taking care so they are never both closed at the same time. Assume that
the voltage across C 2 is initially V 2 and so the initial charge on C 2 is Q2 = C 2 V 2 .
Now S 1 is closed (S 2 is open) so capacitor C 1 is charged to a voltage V 0 . The
charge on C 1 is then Q1 = C 1 V 0 . Next, S 1 is opened and then S 2 is closed. Since the
80 4 More on Capacitors and Inductors

Fig. 4.5 The switching


capacitor arrangement in
(a) can mimic the behavior
of a simple RC circuit, such
as shown in (b)

two capacitors are now in parallel, the total effective capacitance is C 1 + C 2 and the
total net charge on that capacitance is Qtot = Q1 + Q2 = C 1 V 0 + C 2 V 2 . That charge
will rearrange so that the potentials (voltages) across the two capacitors are equal.
The voltage across the effective capacitance is then

Q tot C1 V0 + C2 V2 C1 C2
= = V0 + V2
C1 + C2 C1 + C2 C1 + C2 C1 + C2
C1
= V2 + (V0 − V2 ) = V2 + /\V2 . (4.7)
C1 + C2

After S 2 is again opened, this is the new value for V 2 . If this entire switching
process occurs over a time /\t, then the average change in the charge on C 2 per unit
time is given by

/\Q 2 /\V2 C1 C2 1
= C2 = (V0 − V2 ). (4.8)
/\t /\t C1 + C2 /\t

Compare that to a circuit with a resistor and capacitor, Fig. 4.5b, where at any
time I = dQ2 /dt = (V 0 – V 2 )/R. On time scales long compared to /\t, the switched
capacitor circuit looks the same with an effective resistance

C1 + C2
Re f f = /\t, (4.9)
C1 C2

even though there are no resistors in the circuit. The effective resistance can
be controlled by changing the switching time rather than by changing physical
components.
Such a circuit has many uses, one of which is to measure small capacitance values.
If C 1 is the unknown capacitance then a measurement of the charging time of the
known capacitor C 2 determines R from which you can determine C 1 . The value of
/\t can be adjusted to get convenient values. Another use is illustrated by Van Den
[2], who show how to use a switched capacitor “resistor” as part of a Wheatstone
bridge in order to measure very large resistance values.
4.6 Charging a Capacitor Efficiently 81

4.6 Charging a Capacitor Efficiently

If you charge a capacitor from 0 V to a voltage V 0 using a simple voltage source


and resistor (Fig. 4.6a), the total energy change of the capacitor is 21 C V02 . During the
process, the energy lost in the resistor as heat is found to be the same value. This can
be seen by integrating the power dissipated in the resistor (I 2 R = V 2 /R) over time,

{∞
V2 1
E= 0 e−2t/RC dt = C V 2. (4.10)
R 2 0
0

While this result may seem purely academic, much of the power consumption that
occurs in modern digital electronics arises during what is, effectively, the charging
and discharging of capacitors. If this simple process could be made more efficient, the
electronics could be packed that much tighter without overheating and/or batteries
would last longer.2
One scheme that can reduce the energy loss to almost nothing is to charge the
capacitor in many small steps using gradually increasing battery voltages. It is
straightforward to show that if this is done using n equal steps of size V 0 /n, the
total energy loss is given by E = C V02 /2n. By making n large, the loss is reduced
significantly. Such a procedure may be difficult to implement, but at least it shows
that energy loss is not inevitable and depends on the process. There are many similar
examples of such situations from the field of thermodynamics.
Consider another approach, which is to charge the capacitor using a current source
from time t = 0 (Fig. 4.6b) until it reaches the required final voltage, V 0 . That is,
charge for a time T, such that

{T
1
V0 = I0 dt or T = C V0 /I0 . (4.11)
C
0

The energy lost in the resistor during this time is I02 RT =


C V02 (RC/T ) =C V02 (I0 R/V0 ). The loss can be made very small by taking a
long time compared to RC, which is equivalent to saying that the voltage across the
resistor, I 0 R, is always very small compared to V 0 .3

2 For more detailed considerations, see, for example, [3], and references therein.
3 To learn more about this problem see [4], and related articles.
82 4 More on Capacitors and Inductors

Fig. 4.6 Charging a


capacitor with the circuit
a loses half the energy. Using
the circuit of b and a slow
charge, less energy is lost

4.7 Mutual Inductance and Transformers

Lenz’s law, from Introductory Electricity and Magnetism, says that a time-changing
magnetic field will induce an electromotive force (a voltage) around a closed path,

do
V =− , (4.12)
dt
where o is the magnetic flux through any area bounded by the path and the minus
sign is somewhat symbolic, indicating that the induced voltage is in a direction to
create currents which, in turn, create their own magnetic field in a direction that
opposes the original time-changing magnetic field.
Any circuit with current will produce magnetic fields. If the current changes, those
magnetic fields change, and hence a change in the magnetic flux through the circuit
that created the field in the first place. This gives rise to “self-inductance.” For fixed
geometry and in the absence of magnetic materials, or for small enough changes in
the presence of magnetic materials, the changes in the magnetic flux are proportional
to the changes in the current. That proportionality constant is the self-inductance,
usually signified with the symbol L, and

dI
V = −L . (4.13)
dt
On the other hand, if you have two circuits near each other, the changing magnetic
field from one of them can create a changing flux in the other. Of course, the largest
contributions arise from pieces of the circuit that are designed to create large magnetic
fields (e.g., inductors). Consider two circuits that are magnetically coupled, as shown
in Fig. 4.7. That is, there are two inductors that can “feel” each other’s magnetic field.
If the geometry is fixed, the field felt by circuit 2 on the right is proportional to the
current in circuit 1 on the left, and so.

d I1
V2 = M12 , (4.14)
dt

where the proportionality constant M 12 is called the “mutual inductance.” Similarly,


if there is a changing current in circuit 2 it will create a voltage in circuit 1,
4.7 Mutual Inductance and Transformers 83

Fig. 4.7 Two nearby


inductors can interact to
make a transformer

d I2
V1 = M21 . (4.15)
dt
For any two circuits, no matter how they are constructed, it will always be the
case that M 21 = M 12 ≡ M, a result of the reciprocity theorem. There is still some
arbitrariness about the minus signs. In many practical circuits, those minus signs
are of no consequence. However, when they do matter, there is a “dot convention.”
A small dot is placed on each of the two inductors that allows the minus signs to
be determined accurately. For the moment, this is ignored and hence the resulting
current I 2 may be off by a minus sign. The dot convention will be described later.
Now, to analyze the circuit above, use Kirchhoff’s laws around the two circuits,

V − i ωL 1 I1 − iωM I2 = 0
(4.16)
i ωL 2 I2 + i ωM I1 + R I2 = 0.

These equations can be readily solved to give

R + Z2 ZM
I1 = V I2 = V, (4.17)
R Z1 + Z1 Z2 − Z M Z M R Z1 + Z1 Z2 − Z M Z M

where Z 1 = iωL 1 , Z 2 = iωL 2 , and Z M = iωM.


Now the total power being delivered by the voltage source is VI 1 and the total
power absorbed in the resistor is I22 R. Ideally, those will be the same. Now,

R + Z2 (R + Z 2 )(R Z 1 + Z 1 Z 2 − Z M Z M ) 2
V I1 = V2 = V
R Z1 + Z1 Z2 − Z M Z M (R Z 1 + Z 1 Z 2 − Z M Z M )2
Z 2M R
I22 R = V 2, (4.18)
(R Z 1 + Z 1 Z 2 − Z M Z M )2

and so these will be the same if the numerators are equal, which requires

Z 2M (2R + Z 2 ) = Z 1 (R + Z 2 ), (4.19)
√ √
which can be accomplished if |Z 2 | >> R and Z M = Z 1 Z 2 . That is, M = L 1 L 2 .
Note that M is always taken to be positive. Such a power-conserving transformer is
the so-called “ideal transformer.” Using the result for M and the condition |Z 2 | >> R
in the solutions above, yields
84 4 More on Capacitors and Inductors
/
V L1 V L2
I1 = ; I2 = . (4.20)
R L2 R L1

Most often, this is expressed using the “turns ratio.” That is, if coil 1 is wound
with n1 turns of wire, and coil 2 with n2 , and remembering that all other things being
equal, the inductance is proportional to the number of turns squared, the following
relations can be obtained:
I )2 I ) I )
V n1 n2 n1
= R, I2 R = V, and I2 = I1 . (4.21)
I1 n2 n1 n2

In words, the effective impedance seen by the voltage source is the resistance,
R, “transformed” by the turns ratio squared. The voltage seen by the resistor is
the voltage supplied transformed by the inverse of the turns ratio, and the current
seen by the resistor is the current supplied transformed by the turns ratio. The ideal
transformer changes the ratio of the voltages and currents in a circuit, but not the
power (i.e., the product of voltage and current is preserved). In a real transformer,
some power loss is expected.
Geometrically, the ideal transformer requires that all of the magnetic field lines
created by circuit 1 go through the area bounded by circuit 2 (and vice versa). In
addition, the inductive impedances of the two circuits must be large compared to any
load resistance placed on them.

4.8 The Dot Convention for Transformers

In situations where the relative signs of the currents on the two sides of a transformer
are important, the “dot convention” or “dot rule” is often used. When you see a
transformer (two inductors next to each other) with dots, such as in Fig. 4.8, then
when you assign currents and apply Kirchhoff’s laws, do the following:
If both currents enter or both leave the transformer at the dotted terminal, the sign of
the M terms is the same as that of the corresponding L terms, otherwise the signs of
the two terms are opposite.

Fig. 4.8 Examples of the


use of the dot convention for
transformer windings
4.10 RLC Circuits 85

In the example above, the signs of the M terms were made the same as those of the
L terms somewhat arbitrarily. If dots had been assigned in this case, and the solution
had been done correctly, the inductors would have had one dot on the top and one
on the bottom, as shown on the right of Fig. 4.8. Note that when using dots to get
the minus signs correct, what matters is which direction is defined to be the positive
direction for current—i.e., which way you draw your arrow—and not the direction
the current actually flows. The actual direction will come out in the solution to the
problem.

4.9 Inductive Position Sensors

A simple position sensor can be constructed using coils set up so that the mutual
inductance depends on a position. Two examples are shown in Fig. 4.9. In both
cases, two “sense” coils are used and one excitation coil. The coils are wound so that
the induced voltage measured across the detector, M (not to be confused with the
mutual inductance, M), is zero when the moving portion is centered.

4.10 RLC Circuits

There are a number of combinations of inductors and capacitors that form a tuned
circuit. Of course, any real circuit will also have some resistance, but it is useful to
look at the ideal case as a starting point.
Consider the two simple circuits of Fig. 4.10. For these idealized series and parallel
LC circuits, the (sinusoidal) voltage source sees an effective impedances,

1 1 − ω2 LC
Z series = Z C + Z L = + i ωL = Z parallel
i ωC i ωC
I ) I )
1 1 −1 1 −1 iωL
= + = i ωC + = , (4.22)
ZC ZL i ωL 1 − ω2 LC

Fig. 4.9 Two examples of


an inductive position sensor
86 4 More on Capacitors and Inductors

Fig. 4.10 Basic series and


parallel LC circuits

which have interesting behavior when ω2 LC = 1. In particular, the series impedance


becomes zero and the parallel impedance approaches infinity. Looking at the behavior
for very large and very small frequencies (ω2 LC >> 1 and ω2 LC << 1 respectively),
the limiting behavior summarized in Table 4.1 is obtained. For real circuits replace
“0” with “small” and “∞” with “large” for the impedance values in the table.
Now consider a more realistic series circuit that includes some resistance. In fact,
as an example consider a series circuit with two resistors added (Fig. 4.11): one, R,
that represents a “load” where the signal is delivered (this might be a speaker in a
stereo system, for example) and the other, r, represents all the other resistance in the
circuit, including the resistive losses in the inductor. What is desired is to compute
is the total (time-averaged) power delivered to the load, R, for the circuit, which is
given by the time average of PR = |I |2 R. The current can be found by creating
an effective impedance for the series combination, Z ser , and then I = V 0 /Z ser . The
results are.
( )
1 1 − ω2 LC + i ωC(R + r )
Z ser = r + i ωL + +R=
i ωC iωC
V0 i ωC (R + r )
I = = V0 ( ) ·
Z ser 1 − ω2 LC + i ωC(R + r ) (R + r )
V02 R ω2 C 2 (R + r )2
PR = |I |2 R = I ∗ I R = ·( ) . (4.23)
(R + r ) 1 − ω2 LC 2 + ω2 C 2 (R + r )2

Defining ω0 = 1/ LC, the power dissipated in the load can be written

V02 R (ω/ω0 )2 /Q 2
PR = ·( ) , (4.24)
(R + r ) 1 − (ω/ω0 )2 2 + 1/Q 2

where a “quality factor for the circuit,” Q, has been defined for the series combination
as

Table 4.1 Limiting behavior


Zseries Zparallel
for simple LC circuits
ω2 → 0 → ∞∠ − 90◦ → 0∠ + 90◦
ω2 →∞ → ∞∠ + 90◦ → 0∠ − 90◦
4.10 RLC Circuits 87

Fig. 4.11 A series RLC


circuit

Fig. 4.12 The power


delivered to a load for a RLC
circuit as a function of
frequency for different
quality factors. The higher
the quality factor, the
narrower is the range of
response

ω0 L
Q= . (4.25)
(R + r )

This result for PR is shown graphically in Fig. 4.12 for several different values of
Q. Aside from some name changes, the behavior is the same as the resonant behavior
exhibited by a driven, damped, harmonic oscillator.
Such a circuit can be used as a bandpass filter—that is, only signals with a
frequency within a certain range of frequencies—a “band of frequencies”—are
allowed through to the load while signals with frequencies outside the band are
blocked. As a guide, the width of the “pass band” (the range of frequencies that
“get through”) is roughly /\ω = ω0 /Q. A more precise definition requires a precise
statement for a cut-off between frequencies that “make it through” and those that do
not.
As an exercise, try computing the power delivered to R2 for the circuit in Fig. 4.13.
The result should look very similar to the result above for the series circuit. An
appropriate definition for the “Q of the circuit” here would be Q = (R1 ||R2 )/(ωL).
Low-pass, high-pass, and more complicated band-pass filters can be created using
different arrangements and multiple LC resonant circuits. There are standard tables
available that can be used to help determine values to use. In general, the more
components, the more control one has over the filter properties.
88 4 More on Capacitors and Inductors

Fig. 4.13 A parallel RLC


resonant circuit

4.11 Cable Models

Cables, or more generally “transmission lines,” consist of two (or more) wires in
close proximity. For example, coaxial cables commonly used to connect laboratory
equipment (e.g., with BNC connectors) have an outer jacket acting as one wire, and
an inner conductor as the other. When cables carry electrical signals, both electric and
magnetic fields are involved—hence there must be both capacitance and inductance.
For a cable, these are spread out or “distributed” along the length of the cable and
so it is convenient to describe these quantities “per unit length.” That is, a cable will
have resistance per unit length, r, capacitance per unit length, c, and inductance per
unit length, l.
One of the first models used for cables came from Lord Kelvin as a model for
early trans-Atlantic cables. He included the resistance and capacitance but took the
inductance to be negligible in comparison. That is reasonable only at low enough
frequencies. A discretized model for the cable looks like the circuit of Fig. 4.14a,
where R and C are the resistance and capacitance for some arbitrarily chosen finite
length of the cable, /\x, and so R = r/\x and C = c/\x. Following the procedure used
previously for a similar array made entirely of resistors (see Chap. 2), the first section
is separated and by translational symmetry the remaining (infinite) cable must have
the same effective impedance, as in Fig. 4.14b.
For sinusoidal signals, the equivalent impedance must satisfy
I )
1
Z eq = R + ||Z eq . (4.26)
iωC

It is left as an exercise to show that the solution is

Fig. 4.14 a A simple RC cable model and b an equivalent due to symmetry


4.11 Cable Models 89

R( / )
Z eq = 1 + 1 + 4/(i ω RC) . (4.27)
2
Now to complete the model for a continuous cable, take the limit as /\x becomes
very small. To do that, write R = r/\x, and C = c/\x, where r and c are the resistance
per unit length and the capacitance per unit length respectively. Then take the limit
that /\x becomes infinitesimally small (but r and c stay constant). In that limit the
complex equivalent impedance of the cable is proportional to (iω)−½ , i.e.,
( / ) ( / )
r /\x 4 r 4
Z eq = 1+ 1+ = /\x + /\x + , (4.28)
2 iωr c/\x 2 2 i ωr c
/ /
r 4 r
(iω)− 2 .
1
lim Z eq = = (4.29)
/\x→0 2 i ωr c c

Some care is necessary with the square root of 1/i. It is easy to see that
| I )|
1−i 2 1 − 2i − 1 1
± √ = = −i = , (4.30)
2 2 i

and so
/
1−i 1 r
lim Z eq = ± √ , (4.31)
/\x→0 2 ω1/2 c

and only the solution with the “ + ” sign can be valid for a real cable (why?). This
equivalent impedance has a phase angle of − 45˚° for all frequencies and is, in some
sense, half-way between a resistor (always 0°) and a capacitor (always − 90°).
This model has found use in the study of electronic devices made with Carbon
nanotubes4 indicating that distributed resistance and distributed capacitance are
important even in much smaller devices. It is also an interesting model to motivate
fractional derivatives.
Because it is interesting, a little sidetrack to consider fractional derivations seems
appropriate. The basic idea behind fractional derivatives can be seen by noting that

d 1 iωt d 2 i ωt d α i ωt
e = (i ω)1 eiωt ; e = (i ω)2 eiωt ; e = (i ω)α eiωt , (4.32)
dt 1 dt 2 dt α
and that any function encountered in physics can be written as a Fourier series.
Generalizing this result for terms such as (iω)1/2 eiωt suggests α = ½, that is, a half
derivative. A half derivative applied twice is, of course, the same as a first derivative.
Since the half derivative can be applied to every term in a Fourier series, it can be
applied to any function that can be written as a Fourier series! Of course, one need
not stop at half derivatives, why not make α any real number? You could have the

4 For example, see [5].


90 4 More on Capacitors and Inductors

“square root of two-th” derivative if you set α = 2. Note also that if α < 0 then the
derivative is equivalent to an integral—the inverse operation to a derivative.
Using this notation, the relationship between voltage and current for the infinite cable
model described above can be written in terms of half derivatives:
/ /
r d −1/2 c d 1/2
V (t) = I (t) or I (t) = V (t). (4.33)
c dt −1/2 r dt 1/2

For more information about fractional derivatives, their various definitions and some
of their potential uses, see [6] or do an internet search for “fractional derivatives.”
This ends the sidetrack.

4.11.1 Cable Impedance

For higher frequency applications the inductance in a cable becomes important.


As a starting point to model cables, such as the coax cables commonly used for
experimentation, it is useful to look at the inductance and capacitance and presume
the resistance is negligible. To model such a cable, assumed uniform along its length,
consider the infinite array in Fig. 4.15a, where each segment of length /\x is modeled
using an inductance L = l /\x and capacitance C = c /\x.
To find the equivalent impedance, the infinite array to the right of the first segment
is replaced with its (as yet unknown) equivalent impedance (Fig. 4.15b). Then
( )
Z eq /i ωC iωL ( Z eq +1/iωC )
Z eq = i ωL + i ωC
1
||Z eq = iωC · Z eq +1/i ωC
+ Z eq +1/i ωC
,
( ) iωC
( ) (4.34)
or Z eq 1 + i ωC Z eq = Z eq + iωL 1 + i ωC Z eq ,

so
( ) ( )
Z eq 1 + i ωc/\x Z eq = Z eq + i ωl/\x 1 + i ωc/\x Z eq
(4.35)
i ω Z eq
2
= −ω2 lc/\x Z eq + i ωl,

Fig. 4.15 a A cable model for higher frequencies and b an equivalent due to symmetry
4.11 Cable Models 91


and in the limit /\x → 0, the simple result is that Z eq = ± l/c, , where only the
+ sign is valid here (why?). Note that this equivalent impedance is a real, frequency
independent value. That is, this infinite cable looks like a resistor even though no
resistance was in the model. This is reasonable because when you put power into an
infinite cable, it never comes back, which looks the same as if the power had been
converted to heat in a resistor.
Often a cable will be referred to by its equivalent impedance. For example, a
coaxial cable often used with scientific equipment might be referred to as “50 u
coax cable” and another similar cable (often used for cable TV) is “75 u coax
cable.” Note, however, that the impedance specified is for an infinite length. You
cannot expect to measure that value for any finite length of cable. What will be
measured depends on what is connected at the other end, as will be seen shortly.

4.11.2 Signal Speed in a Cable

To compute the voltage and current inside the cable, start by considering a spot in the
cable, which will be labeled x = 0, where there is a sinusoidal voltage with angular
frequency, ω. Referring to Fig. 4.16 and using the discrete model, the voltage a small
distance /\x away can be computed using the voltage divider equation,

(Z eq ||1/i ωC) Z eq
V (/\x) = V (0) = ( ) V (0)
i ωL + (Z eq ||1/iωC) Z eq 1 − ω2 LC + iωL
Z eq
= ( ( )) V (0). (4.36)
Z eq 1 − ω2 lc /\x 2 + i ω(l/\x)

To take the limit as /\x → 0, use the approximation

1
≈ 1 − ε, (4.37)
1+ε

valid for ε << 1, and drop the /\x 2 terms compared to the /\x terms to get
I )
i ωl
V (/\x) = 1 − /\x V (0). (4.38)
Z eq

This can be rearranged to get

Fig. 4.16 The discretized


model for the transmission
line used to derive the signal
along the cable
92 4 More on Capacitors and Inductors

V (/\x) − V (0) i ωl
= . (4.39)
/\x Z eq

Taking the limit /\x → 0, including the limiting value for Z eq found previously,
and using the fact that the location of x = 0 was arbitrary, this equation becomes

dV ( √ )
= − i ω lc V, (4.40)
dx

that has as its solution V (x) = V 0 e−ikx , where V 0 is a constant and k = ω lc.
That is, the
√solution is also sinusoidal with position
√ along the cable, with wavelength
λ = 2π/ ω2 lc. √ Since f = 2π ω, f λ = 1/ lc. Hence, the wave is traveling with
wave speed 1/ lc. For typical cables, the wave speed is a good fraction of the speed
of light.

4.11.3 Impedance of Finite Cables

Of course, real cables are not infinite. Now compute the effective impedance seen
at one end of a cable when an impedance Z, a “load,” is attached at the other end,
illustrated in Fig. 4.17. The characteristic cable impedance (Z eq above, in the limit
/\x → 0) is taken to be Z 0 .
For signals at frequency ω, the voltage and current in the finite cable can be
described as the superposition of two traveling waves, one traveling toward the load
and the other away from the load. That is

V (x) = VR ei(ωt−kx) + VL ei (ωt+kx) . (4.41)

The current will then be given by

VR i(ωt−kx) VL i(ωt+kx)
I (x) = e − e , (4.42)
Z0 Z0

Fig. 4.17 An impedance at one end of a cable may appear to be completely different if measured
at the other end
4.11 Cable Models 93

where the sign convention used has I(x) > 0 corresponding to a total current traveling
to the right and a minus sign is added for current traveling to the left. Note that in
general, V (x)/I(x) /= Z 0 .
The two unknown values, V R and V L , can be determined by the conditions

V (0) = VS eiωt (4.43)

and
V (l)
= Z. (4.44)
I (l)

Applying the condition at x = 0 gives

VS = V R + V L . (4.45)

The condition at x = l gives

VR e−ikl + VL e+ikl
Z = Z0 , (4.46)
VR e−ikl − VL e+ikl

or, solving this for V L ,

Z − Z0
VL = VR e2ikl . (4.47)
Z + Z0

Putting that result back into the condition at x = 0 gives


I )
Z − Z 0 2ikl −1
V R = VS 1 + e , (4.48)
Z + Z0

and then
I )
VS Z 0 cos(kl) − i Z sin(kl)
I (0) = (VR − VL )/Z 0 = (2VR − VS )/Z 0 = .
Z0 Z cos(kl) − i Z 0 sin(kl)
(4.49)

The “effective impedance,” Z eff , seen by the signal source is then


I ) I )
VS Z cos(kl) − i Z 0 sin(kl) Z − i Z 0 tan(kl)
Ze f f = = Z0 = Z0 . (4.50)
I (0) Z 0 cos(kl) − i Z sin(kl) Z 0 − i Z tan(kl)

In addition to computerized solutions, a special graphic known as a Smith chart


may be useful to those for whom such computations occur often. If Z = Z 0 , then
94 4 More on Capacitors and Inductors

Z eff = Z 0 and the cable is “impedance matched” to the load.5 One other result that is
easy to show is that if kl = nπ, where n is any integer (that is, the transmission line
length is a multiple of a half wavelength, λ/2), then Z eff = Z regardless of the value
of Z 0 .
Of interest for some applications are transmission lines (cables) where kl = (n +
½)π, the shortest being a quarter wavelength (n = 0). It is easy to show that for a
quarter-wavelength transmission line

Z 02
Ze f f = , (4.51)
Z
and in particular, if Z is a short circuit, Z eff corresponds to an open circuit, and if
Z is an open circuit, Z eff is that of a short circuit. In addition, if Z corresponds to
a capacitance, Z eff looks like an inductance and vice versa. The effect of the cable
length is very far from negligible in this case.
In practice, these results show that if the cable is very short compared to a wave-
length (i.e., the signal’s wavelength within the cable) then the effects of the cable
can be neglected. On the other hand, once the cable becomes comparable to a fair
fraction of the wavelength, the cable can have a huge effect. Remember that the speed
of light is roughly 1 foot (25 cm) per nanosecond, and modern electronics can have
clock speeds over 1 GHz (that is, less than 1 ns per clock pulse), so even a relatively
short cable may need to be treated carefully.
For most commercially available cables the wave speed is about 0.5–0.8 times
the speed of light in a vacuum. Hence, minimum cable lengths that may become of
concern can be estimated from those values. For frequencies below 1 MHz, a cable
10 m long will not be much different from a short cable, however at 100 MHz, even
a 10 cm cable may be significant.

4.12 Capacitor and Inductor Labels

It is common practice that numerical identifiers on capacitors are usually either in


pF or μF. If the value does not include units, then if the value shown is below 1.0, it
is (probably) μF and if the value is larger than 1, it is (probably) pF. Increasingly nF
and mF are also being used, though generally not without units displayed. Hence, a
capacitor labeled 0.1 would probably be 0.1 μF and a capacitor labeled 33 would be
33 pF. For values in pF, a third number is often used for a power of ten. For example,
a capacitor labeled “104” would be 10 × 104 pF, which is the same as 105 pF =
0.1 μF, and a capacitor labeled “561” would be 560 pF. If the third digit is a zero,
however, it is probably a third digit. Thus, a label 150 would (probably) be 150 pF,

5 On some equipment, the input/output connectors may be labeled with a resistance value such as
“50 u” or “1 Mu.” These values are the equivalent input/output impedance of the device and are
the load at the end of the cable connected to the device.
4.13 Duality 95

the same as a label 151. As for resistors, an “R” is sometimes used as a decimal point
for smaller capacitors. For example, 4R7 would be 4.7 pF.
On many capacitors, a single uppercase letter follows the value of the capacitor.
This is usually not a multiplier but a tolerance code, a temperature coefficient, or
some other identifier.
Numerical identifiers on inductors are highly variable and depend much more
strongly on the manufacturer and intended use than is seen for resistors and capacitors.

4.13 Duality

In electronics, and more generally in electricity and magnetism, there is a concept


called duality. For electronics, if two devices or circuits are related by the interchange
of voltage and current, then they are called duals. Voltage sources and current sources
are obvious duals. Ohm’s law is usually written V = IZ, where Z is the (complex)
impedance. Defining the (complex) admittance, Y ≡ 1/Z, Ohm’s law can equally
well be written I = YV. This has the same form as the previous version, but the roles
of current and voltage have been switched. Hence, impedance and admittance are
duals.
Inductors and capacitors are also duals. The basic relationships are

d VC d IL
IC = C and VL = −L , (4.52)
dt dt
which are of the same form except the roles of current and voltage have been
switched. Likewise, series resistors (that make a voltage divider) and parallel resis-
tors (that make a current divider) are duals. Defining the conductance, G ≡ 1/R, the
voltage divider written for series resistors and the current divider written for parallel
conductances are6
Rk Gk
Vk = Vo E and Ik = I0 E , (4.53)
all Ri all Gi

which are of the same form except the roles of current and voltage have been switched.
Entire circuits can have a dual circuit—one with equations to solve that are the
same except that voltage and current have been swapped. Duals will not play a
significant role in this text beyond this brief introduction to them. It should be noted,
however, that sometimes when solving a circuit, or perhaps for understanding a
circuit, the task might be made easier by considering a dual.

6 These dividers were first introduced in Chap. 1.


96 4 More on Capacitors and Inductors

4.14 Problems

1. Consider the two circuits in Fig. 4.P1 for the special case when R = ωL = 1/
ωC, where ω is the angular frequency of the voltage source. Hence Z L = iR and
Z C = − iR. (a) Compute I 0 and V out for both circuits. (b) Now compute I 0 for
both circuits for the case where the output has been shorted to ground (so V out
= 0).

Fig. 4.P1 Problem 1

2. The bridge circuit in Fig. 4.P2 is to be used to determine L and its resistance
using the parallel model. A known capacitance, C, and resistance R2 , are varied
to find a null condition across R5 . Once the null condition is found, what is the
value of the unknown inductance and its parallel resistance, RLP , in terms of R1 ,
C, and R2 ?

Fig. 4.P2 Problem 2

3. The switching circuit in Fig. 4.P3 has two pairs of switches. The switches operate
together within each pair, as indicated by the dashed line. That is, the pair of
switches S1 and S2 are either both on or both off, and similarly for the pair S3
and S4 . In operation, the two pairs are switched back and forth such that only
one pair is on at any given time. If this switching is done very rapidly compared
to RL C 2 , what is V out ?
4.14 Problems 97

Fig. 4.P3 Problem 3

4. One common coaxial cable used with scientific instrumentation is referred to as


“RG-58.” While there are several varieties, the cable impedance will be close to
50u and will have a capacitance per unit length of about 100 pF/m. As a fraction
of the speed of light, how fast do electric signals propagate in such a cable?
5. Assuming a force that is sinusoidal in time with angular frequency ω, and using
imaginary numbers as was done in Chap. 3 for capacitors and inductors, derive
the relationship between force and velocity corresponding to Newton’s second
law (F = ma) and between the force and velocity for an ideal spring (F = –kx).
For what value of ω will those forces be equal?
6. A student solves a particular circuit that includes some resistors, a capacitor, and
an inductor. One of the following three results is correct. Which one is it and
why?
(a) Vout = R1 R2
V
ω2 LC−R1 R2 in
ω R1 C1 +1
(b) Vout = ω2 LC−1
Vin
1 (R1 +R2 )C
(c) Vout = ω (LC−1)
2 Vin
7. (Challenge Problem) A capacitor is charged through an inductor, as shown in
Fig. 4.P7. The switch is closed at t = 0 at which time the current begins to
rise and then to oscillate. When the current next becomes zero, the switch is
re-opened. After this process, what is the voltage across the capacitor? Assume
the quality factor, Q = ωL/r, is much larger than 1. Show that the energy lost
due to the resistance is proportional to 1/Q, and hence can be much less than the
energy stored in the capacitor.

Fig. 4.P7 Problem 7


98 4 More on Capacitors and Inductors

References

1. R.V. Jones, J.C.S. Richards, The design and some applications of sensitive capacitance
micrometers. J. Phys. E 6, 589–600 (1973)
2. J.A. Van Den Akker, G.M. Webb, A method for measuring high resistance. Rev. Sci. Instruments
7, 44–46 (1936)
3. S. Paul, A.M. Schlaffer, J.A. Nossek, Optimal charging of capacitors. IEEE Trans. Circ. Syst. I
47, 1009 (2000)
4. F. Heinrich, Entropy change when charging a capacitor. Am. J. Phys. 54, 742–744 (1986)
5. G. Esen et al., Transmission line impedance of carbon nanotube thin films for chemical sensing.
Appl. Phys. Lett. 90, 123510 (2007)
6. I.M. Sokolov, J. Klafter, A. Blumen, Fractional kinetics. Phys. Today 55(11), 48–55 (2002)
Chapter 5
The Laplace Transform

The Laplace transform is a linear operation that can be performed on functions of a


single variable. For electronics, the transformation is taken from the time domain (the
single variable is t) to the “s-domain” where s is a generalized complex frequency.
The transform is useful for solving systems of linear differential equations such as
those found for some electronic circuits. An understanding of the transform can lead
to a description of circuit behavior expressed in terms of poles and zeros in the
s-domain.
As a word of warning, this chapter is somewhat mathematically intense. The
mathematics is not particularly advanced, at least for those who have made it through
calculus, however there is a lot of it.
A mathematical transform is a way of rearranging all the information contained
within a function. An invertible transform loses no information so you can get back
to where you started. It is like rearranging the books on a bookshelf—an arrangement
of the books based on alphabetic order by title can be transformed to an arrangement
based on alphabetic order by author without losing any books. The problem of finding
all the books by a certain author is certainly easier if the books are in order by author
rather than by title. The mathematics is much more abstract, but that is, in essence,
what is happening here.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 99


B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_5
100 5 The Laplace Transform

5.1 The Transform

One form of the Laplace transform, F(s), of a function f (t), is given by1
{ ∞
L( f (t)) = F(s) = f (t)e−st dt, (5.1)
0+

where s is a complex number and “0+” means the lower limit should be interpreted to
be “just after t = 0.” For the discussion here, t is time and the function is transformed
to the “s-domain.” The inverse transform can be written as a complex integral along
a particular path in the complex domain and will not be reproduced (or used) here.
It should be noted that the Fourier transform can be considered to be the special
case of the Laplace transform where s is restricted to be a purely imaginary number.
That is, in general write s = σ + iω and for the Fourier transform include only
the special case where σ = 0. Note also that in general the frequency, ω, can be
any real number including both positive and negative—that is, for this use, negative
frequencies are taken as distinct from positive frequencies.
The Laplace transform is invertible—there is uniqueness so that there is a one-to-
one correspondence between a function and the function that is its transform. They
appear in pairs. To execute the appropriate integral for the inverse is not always
a simple thing to do. Fortunately, for electronics applications, one rarely needs to
compute these integrals. Instead, a table of Laplace Transform Pairs is used. One
simply looks up the answers. A selection of such pairs is included in Table 5.1.
An advantage of the Laplace transform is that operations and functions such
as derivatives, delta functions, derivatives of delta functions, etc., become nicely
behaved in the s-domain making solutions, in some sense, very easy. The major
work becomes transforming the problem and then transforming back after the easy
solution is found.
Using transforms of various kinds is a standard practice in Physics. A problem is
transformed into another one that is easier to solve, then transformed back. One of
the simplest ways this is accomplished is to transform from one coordinate system
to another. Indeed, the Laplace transform is a generalization of that idea where the
t-coordinate axis is transformed into an s-coordinate axis.
Since the transform is an integral, which is a linear operation, the transform itself
is linear. That is, for any two constants a and b and any two functions f (t) and g(t):

L(a f (t) + bg(t)) = aL( f (t)) + bL(g(t)) . (5.2)

This linearity is important for finding solutions by reverse look-up.


The use of the Laplace transform will be illustrated using the concrete examples
that follow. These examples from electronics use a look-up table to do the transforms

1 The Laplace Transform is named after the French mathematician Pierre Simon, Marquis de
Laplace, (1749–1827), who is credited with its development as well as with many other discoveries
in mathematics, physics, and astronomy.
5.1 The Transform 101

Table 5.1 Some laplace transform pairs


f (t) F(s)
1 1 1/s
2 t 1/s 2
3 e−at 1/(s + a)
4 te−at 1/(s + a)2
5 t n−1 −at
(n−1)! e 1/(s + a)n = (s + a)−n
( )
6 sin ωt ω/ s 2 + ω2
( )
7 cos ωt s/ s 2 + ω2
| |
8 e−at sin ωt ω/ (s + a)2 + ω2
| |
9 e−at cos ωt (s + a)/ (s + a)2 + ω2

10 c2 + d 2 e−at cos(ωt − tan−1 (d/c)) c(s+a)+dω
(s+a)2 +ω2
( )
11 sinh ωt ω/ s 2 − ω2
( )
12 cosh ωt s/ s 2 − ω2
Special transforms
( )
13 df
dt s F(s) − f 0+
14 {t F(s)/s
f (τ )dτ
0

15 c1 f 1 (t) + c2 f 2 (t) c1 F 1 (s) + c2 F 2 (s)


16 {t F 1 (s)F 2 (s)
f 1 (t) f 2 (t − τ )dτ
0

17 δ(t − T ) (delta function, T > 0) e−sT


dn
18 dt n (δ(t)) (n-th derivative of delta function) sn
19 δ (−n) (t) (n-th integral of delta function) s −n
Comments on the Transform Table
(A) The transform variable, s, is complex. The Fourier transform is the special case of this more
general transform where s is restricted to be pure imaginary. Thus, such a table can also be used for
Fourier transforms
(B) In the table above, a, c, and d are considered real constants, though many of the transforms are
still valid if they are complex or pure imaginary. Consult a more general table for those cases
(C) 1–4 are special cases of 5
(D) 1, 3, and 6–9 are special cases of 10
(E) 6 and 7 are related to 11 and 12. If you replace ω with iω in one pair, you generate the other
(F) 15 is a statement of the linearity of the transform
(G) 14 is a special case of 16 (where one of the functions is constant)
(H) 16 is known as the “convolution theorem”
(I) 18 and 19 are particularly useful if you have an equation with a delta-function in it and need
to take a derivative. This occurs for some idealized models of real systems (e.g., In E&M, current
through a thin wire is often modeled using a delta function for the current density)
102 5 The Laplace Transform

and inverse transforms. The difficulty will be putting the inverse problem into a form
that appears in the look-up table.

5.2 Laplace Transform Example 1

Consider the circuit in Fig. 5.1, where the switch is open for t < 0 and closed for t
> 0. The previous methods (simple L/R time constants and/or complex impedances)
will not work for this circuit because it has a sinusoidal source and a switch.
Using Kirchhoff’s voltage law around the loop (after the switch is closed),

dI
V0 sin(ω0 t) − L − I R = 0, (5.3)
dt
and the objective is to find I(t). Note that here ω0 is the frequency of the source and
is a constant, and is not the same as the variable ω, which is the imaginary part of s.
Now transform each term in the equation (e.g., using a look-up table) to get
ω0
V0 − L(s I (s) − I (0)) − I(s)R = 0, (5.4)
s 2 + ω02

which has transformed the immediate problem to that of finding I(s). Note that I(0)
refers to the current at t = 0 and for this problem I(0) = 0 (the inductor keeps the
current constant, as best it can, and before the switch was closed the current was
zero). Solving for the current in the s-domain is now straightforward algebra, and

V0 ω0 1 V0 ω0
I (s) = = (5.5)
L s + ω0 s + R/L
2 2 L (s + i ω0 )(s − i ω0 )(s + R/L)

A strategy being used is that, if possible, the result is written using products of
terms where the coefficient in front of each “sn ” term is one. The second part of the
strategy is to try to write it as a product involving terms such as (sn + a), where a is
some (possibly complex, possibly zero) constant. For this example, n = 2 or n = 1
for all the terms in the denominator. This is done because such terms appear in the
look-up table. Also, this makes the denominator look like the denominator obtained
when several fractions are added using a common denominator. Now the hardest part
of the problem is to transform back to find I(t).

Fig. 5.1 A simple LR


circuit with a switch used as
an example
5.2 Laplace Transform Example 1 103

The expression on the right above does not appear in the look-up table. One way
to express it using terms that do show up is to use the method of partial fractions,
illustrated below.

5.2.1 Method I

Noting that the individual terms in the denominator are in the table, guess a solution
using unknown constant values A, B, and D, and then require that2

A B D ω0
+ + = . (5.6)
s + i ω0 s − i ω0 s + R/L (s + i ω0 )(s − i ω0 )(s + R/L)

The constant out front (V 0 /L) is dropped for now, and will simply be reintroduced
at the end. The problem is to find constants A, B, and D that make this guess work.
Creating a common denominator and adding, the left side becomes

A(s − iω0 )(s + R/L) + B(s + iω0 )(s + R/L) + D(s + iω0 )(s − i ω0 )
(left side) = .
(s + iω0 )(s − iω0 )(s + R/L)
(5.7)

This will equal the right side only if the numerators are the same for all s. To get
the numerators to be the same for all s, the coefficients for each power of s must be
equal. In this example, the only power of s in the numerator on the right side of (5.6)
that has a non-zero coefficient is s0 (i.e., the constant term, ω0 = ω0 s 0 ). Now expand
the numerator on the left side and make the coefficient for each power of s have the
same coefficient as on the right. That is

s 2 terms: (A + B + D) = 0
s 1 terms: A(R/L − i ω0 ) + B(R/L + iω0 ) = 0 (5.8)
s 0 terms: A(−i ω0 R/L) + B(i ω0 R/L) + Dω02 = ω0 ,

which now need to be solved for A, B, and D. If a unique solution cannot be found,
the initial guess was bad.
For this example, after some algebra, and defining τ = L/R,
( )
τ 1 + i ω0 τ 1 ω02 τ 2
A=− ; B = A∗ ; D= . (5.9)
2i 1 + ω02 τ 2 ω0 1 + ω02 τ 2

Note that A, B, and D are now all constants with known values. Now write
(reintroducing the constant out front, V 0 /L),

2 “C” is skipped here so that there is no confusion with any capacitance values.
104 5 The Laplace Transform
( )
V0 A B D
I(s) = + + (5.10)
L s + i ω0 s − i ω0 s + R/L

and all these terms are in the form 1/(s + a) which is in the Laplace transform tables.
When transformed back such terms result in exponentials. Hence, using the table

I (t) = V0
L (
(Ae−i ω0 t)+ Bei ω0 t + De−t/τ )
V0 τ
(1 )
= 1
L ( 1+ω02 τ 2 2i (1
− i ω0 τ )ei ω0 t − 2i1 (1
+ i ω0 τ )e−iω0 t + ω0 τ e−t/τ
( ) ) (5.11)
√ ω0 −t/τ
= V0
R
1
sin(ω0 t − ϕ) + 1+ω2 τ 2 e ,
1+ω02 τ 2 0

where φ = tan−1 (ω0 τ ). Note that for long times the solution is the same as obtained
from the usual AC analysis using complex impedances—eventually one cannot tell
that a switch was closed a long time ago. The term on the far right is a transient that
is present just after the switch is closed and the transient will persist for times of
order τ = L/R. Hence in this case “a long time” means long compared to L/R.

5.2.2 Method II

Multiplying the first two terms in the denominator of I(s),

V0 ω0
I (s) = ( ) , (5.12)
L s 2 + ω02 (s + R/L)

each of the terms in the denominator appears in the table separately. Hence, using
constants A, B, and D (none of which are assumed to be the same values as found in
Method I), try

As + B D ω0
( )+ =( 2 ) , (5.13)
s + ω0
2 2 (s + R/L) s + ω0 (s + R/L)
2

noting that the numerators in our guess should be one power less than their corre-
sponding denominator. Creating a common denominator and adding on the left
side

As 2 + As R/L + Bs + B R/L + Ds 2 + Dω02 ω0


( ) =( 2 ) , (5.14)
s + ω0 (s + R/L)
2 2
s + ω0 (s + R/L)
2

and once again, the coefficients in front of each power of s must match:
5.3 Laplace Transform Example 2 105

s 2 terms: A + D = 0
s 1 terms: A R/L + B = 0 (5.15)
s 0 terms: B R/L + Dω02 = ω0 ,

which yields

L ω0
A = −B = −D = 2 . (5.16)
R ω0 + R 2 L 2

Once again defining τ = L/R and putting these known constants back into the initial
guess,
( )( )
V0 ω02 τ 2 −s 1/τ 1
I (s) = + + , (5.17)
ω0 L 1 + ω02 τ 2 s 2 + ω02 s 2 + ω02 s + 1/τ

which can easily be inverted using Table 5.1 (entries 3, 6, and 7) to give
( )
V0 1 ( )
I (t) = −ω0 τ cos ω0 t + sin ω0 t + ω0 τ e−t/τ , (5.18)
R 1 + ω0 τ
2 2

which can be made to match the form of the final result from Method I through use
of the identity3
/
p cos z + q sin z = p 2 + q 2 sin(z + φ); φ = tan−1 ( p/q). (5.19)

Note also that while Kirchhoff’s voltage law produced a differential equation, the
Laplace Transform eliminated the derivative making the problem easy to solve in
the s-domain using simple algebra. The Laplace transform may also be a useful tool
to solve many other differential equations, including equations with delta functions,
derivatives of delta functions, non-linear terms, and possibly some other pathological
terms.

5.3 Laplace Transform Example 2

Consider the circuit of Fig. 5.2, where the switch has been open for a long time and
is closed at t = 0 (the capacitor is initially uncharged). Now solve for the currents
in the circuit for t > 0. This circuit can be solved by finding the appropriate time
constant, writing an exponential solution, and applying the appropriate initial and
“long time” values. On the other hand, if the voltage source were sinusoidal, or even

3 As was the case for complex phase angles, the arctangent must be used with care.
106 5 The Laplace Transform

Fig. 5.2 A two-loop RC


circuit used as an example

more complicated, such a solution is not valid. Consider solving this with the Laplace
transform.
For t > 0 Kirchhoff’s voltage law around the two loops shown yields two equations,

V0 − I1 R1 − (I1 − I2 )R2 = 0
{t (5.20)
(I1 − I2 )R2 − I2 R3 − C1 I2 dt = 0,
0

where the integral yields the net charge accumulated on the capacitor since t = 0.
Since the charge started at zero for this example, this integral also gives the total
charge on the capacitor.
Applying the Laplace transform to both sides of both equations, indicating the
transformed function with bold, and with a little rearrangement,

I 1 (R1 + R2 ) − I 2 R2 = V0 /s
(5.21)
I 1 R2 − I 2 (R2 + R3 + 1/(sC)) = 0.

Solving the second of these to get I2 in terms of I1 , and putting that result back into
the first yields solutions for I1 and I2 :

I 2 = R2 +R3R+1/(sC)
2
I1
( ) (5.22)
I 1 R1 + R2 − R2 R2 +R3R+1/(sC)
2
= V0
s
.

Now, in an attempt to simplify the notation, define the constants

α = (R2 + R3 )C; β = (R1 R2 + R1 R3 + R2 R3 )C; γ = (R1 + R2 )/β, (5.23)

and then
| |
α s + 1/α α 1 1/α
I 1 = V0 = V0 +
β s(s + γ ) β s+γ s(s + γ )
s R2 C/α R2 C 1
I2 = I 1 = V0 . (5.24)
s + 1/α β s+γ
5.4 Laplace Transform Example 3 107

Now these currents must be transformed back to the time domain. From the look-up
table:
1 1 1
→ 1; → e−γ t ; → e−t/α . (5.25)
s s+γ s + 1/α

The products in the denominator for I1 suggests that the method of partial fractions
might work. That is, for I1 find constants A and B such that

A B A(s + γ ) + Bs 1/α
+ = = , (5.26)
s s+γ s(s + γ ) s(s + γ )

which will be true if the coefficients in front of each power of s in the numerator is
the same on both sides. That is,

A + B = 0 and Aγ = 1/α, (5.27)

which gives

1 −1
A= , B= , (5.28)
αγ αγ

so the solution can be written


( )
α −γ t 1 ( )
I1 = V0 e + 1 − e−γ t
β αγ
R2 C −γ t
I2 = V0 e . (5.29)
β

Substituting in values for α, β, and γ , yields the final result.

5.4 Laplace Transform Example 3

Consider the circuit in Fig. 5.3 where initially (at t = 0) the current is zero and the
charge on the capacitor is Q0 with positive charge as defined in the schematic. This
problem is analogous to a damped harmonic oscillator starting at rest that is struck
at t = 0.
Writing Kirchhoff’s voltage law around this loop,

Q dI
+ IR+ L = 0, (5.30)
C dt
where the charge on the capacitor at time t (t > 0) is given by
108 5 The Laplace Transform

Fig. 5.3 An RLC circuit


used as an example

{t
Q = Q0 + I dt. (5.31)
0

Using the Laplace transform on both sides of the equation,

Q0 I (s)
+ + R I (s) + L(s I (s) + I (0)) = 0, (5.32)
sC sC
and solving for I(s), using I(t = 0) = 0, as specified above, yields

1 Q0 1
I(s) = Q 0 = . (5.33)
sC(s L + R + 1/(sC)) LC (s 2 + s R/L + 1/(LC))

For convenience, define ω0 = 1/(LC) and 2γ = R/L, so this can be written

1
I (s) = ω02 Q 0 ( 2 ). (5.34)
s + 2γ s + ω02

Now the table does not include a generic quadratic but it does include a few entries
that are close. Instead, use the quadratic formula to find the zeros, and then write
the denominator as a product of two first order terms. The zeros of the quadratic are
given by
( ) ( / )
/
1 γ 2
s= −2γ ± 4γ 2 − 4ω02 = −γ ± i ω0 1 + 2 , (5.35)
2 ω0
/
and defining ω' = ω0 1 − (γ /ω0 )2 , gives

1 1 1
= · . (5.36)
s 2 + 2γ s + ω02 s + (γ + i ω' ) s + (γ − i ω' )

Using partial fractions with the constants A and B,

A B 1 1
'
+ '
= · , (5.37)
s + (γ + i ω ) s + (γ − i ω ) s + (γ + i ω ) s + (γ − i ω' )
'
5.4 Laplace Transform Example 3 109

and creating a common denominator on the left, the two will be equal if A and B are
chosen so that the numerators are equal. That is, for all s, A and B must satisfy
( ) ( )
A s + γ − i ω' + B s + γ + i ω' = 1, (5.38)

and as before, the solution can be found by equating all powers of s. That is

s 2 and higher → none.


s 1 → A(+ B = 0) ( )
(5.39)
s 0 → A γ − i ω' + B γ + i ω' = 1,

which is easily solved to give

i
A= ; B = −A. (5.40)
2ω'

Hence,
( )( )
1 1 1
I (s) = ω02 Q 0 − , (5.41)
2ω' s + γ + iω' s + γ − i ω'

which can be readily transformed using the table to give


( ω )2 ( )
1 ( −(γ −i ω' ) )
− e−(γ +iω )
'
' 0
I (t) = ω Q0 e
ω' 2i
( ω )2
= ω' Q 0 e−γ t sin ω' t.
0
(5.42)
ω
The solution looks like a damped sinusoid, as should be expected. However, one
must be careful, as this conclusion is actually only valid if ω' is real and non-zero (the
“underdamped case”), which was implicitly assumed but which is not necessarily
true. The case where γ /ω0 > 1 (the “overdamped case”) results in ω' , as defined
above, being pure imaginary. That case is easily treated using identities /for the sine of
a complex value, rather than starting over. That is, redefining ω' = ω0 (γ /ω0 )2 − 1,
so that it is real and positive, the solution becomes
( ω )2 ( ) ( ω )2 ( )
I (t) = i ω' ' Q 0 e−γ t sin −i ω' t = ω' ' Q 0 e−γ t sinh ω' t
0 0
ω ω
ω' ( ω0 )2 ( −(γ −ω' )t −(γ +ω' )t
)
= Q 0 e − e . (5.43)
2 ω'

As a function of time, this solution grows to a maximum and then decays to zero.
Consult a good set of math tables for the trigonometric identities used for the complex
arguments.
110 5 The Laplace Transform

Now if γ /ω0 = 1, or equivalently ω' = 0, then the solution above is invalid because
during the solution both sides of the equation were multiplied by zero. Rewriting
I(s) for this special case (known as “critical damping”) one gets

1
I (s) = ω02 Q 0 . (5.44)
(s + γ )2

which is found on line 4 of Table 5.1 and is easily inverted to give

I (t) = ω02 Q 0 t e−γ t , (5.45)

that also starts at zero, grows, and then decays to zero. Though it may not be obvious,
this solution can also be obtained by taking the limit ω' → 0 using either of the
previous solutions.

5.5 Comment on Partial Fractions

Whenever there is a polynomial in the denominator of degree n, it can be factored,


as above, or the guess can include a polynomial in the numerator of degree n − 1.
For example, to invert

1
F(s) = ( 2 ), (5.46)
s s + b2

try

A Bs + D 1
+ 2 = (5.47)
s s + b2 s(s 2 + b2 )

with A, B, and D unknown constants to be determined. Note that the total number
of constants necessary will be the same no matter how the fraction is divided. Using
this method only helps if the remaining expressions can be found in the Laplace
transform look-up table. Example 1, Method II, uses this method, and Example 3
above could have been done using this method.
Whenever there is a degeneracy in the denominator, that is, the same quantity to
a power larger than 1, then do the partial fractions including all powers up to and
including the one in the denominator. That is, to use partial fractions to invert

1
F(s) = , (5.48)
s(s + b)3

that has the root −b appearing three times, start by using four unknown constants
and write
5.6 Poles and Zeros 111

A B C D 1
+ + + = . (5.49)
s (s + b) (s + b) 2
(s + b) 3
s(s + b)3

This section on the partial fraction method was for review and should not be
considered a complete treatment by any means. For more information about the
method of partial fractions, consult an appropriate mathematics text.

5.6 Poles and Zeros

In general, the currents in a linear circuit can often be expressed in the s-domain in
the form
(s − z 1 )(s − z 2 ) · · · (s − z n )
I (s) = A (5.50)
(s − p1 )(s − p2 ) · · · (s − pm )

where zi are “zeros” of the response (where the numerator is zero) and pi are “poles”
of the response (where the denominator is zero). Both zi and pi may be complex
values, including pure real and pure imaginary values.
In the Laplace Transform Example 1 (see Method I) above there were no zeros
and three poles (iω, − iω, and − R/L). In Example 2, I 1 has two poles and one zero
and I 2 has one pole, and in Example 3 there were two complex poles.
More generally, one may wish to consider the response (a current or voltage) at
one position in a circuit due to an excitation (a current or voltage source) at some other
position in the circuit. For linear circuits, the ratio of the response to the excitation,
known as the transfer function, can always be expressed in the form shown above. One
place this occurs is in “input–output” devices (which includes filters and amplifiers)
that look like the simplified circuit of Fig. 5.4.
Hence, if the circuit is linear, V out /V in can be put in the form shown above (for
the s-domain) and can be described using poles and zeros. If the poles and zeros are
specified, the behavior of the circuit, aside from a multiplicative constant, has also
been specified. Thus, the difficult part of the problem, the inverse, can be avoided if

Fig. 5.4 A general


representation of a circuit
with an input and output
112 5 The Laplace Transform

one has an understanding of poles and zeros. The main reason for considering the
Laplace transform here is, in fact, to get some of that understanding.
In many cases, the behavior of even a very complicated circuit, at least over the
range where a circuit will actually be used, will be dominated by just a few poles
and/or zeros. Not all of the poles and zeros need to be known. Hence the use of these
dominant poles and zeros is a convenient way to specify the important behavior of
many circuits, or more generally, any system with behavior that can be described
using linear differential equations.

5.7 Problems

1. Show that if X(s) is the Laplace transform of x(t), and y(t) = x(t)e−αt , then Y(s)
= X(s + α).
2. Show that if x(t) is a square wave with period T, such that
{ ( )
1, (nT ≤ t) < n + 21 T
x(t) =
−1, n + 21 T ≤ t < (n + 1)T

where n is any integer, then the Laplace transform of x(t) is given by

1 1 − e−sT /2 1
X(s) = · = tanh(sT /4).
s 1 + e−sT /2 s

Recall that the sum for an infinite geometric series is


( 2 )
a + ar + ar 2 + ar 3 + · · · = a/(1 − r ) |r | < 1 .

3. Using the result above for the square wave, derive the Laplace transform for a
triangle wave by noting that a triangle wave can be written as a time integral of
a square wave.
4. For the circuit of Fig. 5.P4, the source operates at a frequency of f 0 = 50 Hz and
the switch is closed at t = 0. Use the Laplace transform method to solve for the
current through the inductor in this circuit when t > 0.

Fig. 5.P4 Problem 4

5. In the process of solving for the current in a circuit using the Laplace transform
method, a student, Gunner Dufein, found that
5.7 Problems 113

3s
I(s) = mA.
(s + 1)(s − 2)

Identify the poles and zeros associated with the circuit. What is this current
as a function of time?
Chapter 6
Diodes

This chapter considers diodes—in particular, semiconductor diodes. Diodes are non-
linear devices that pass current much more easily in one direction than in the opposite
direction. The emphasis here is to learn techniques of simple analysis for circuits
that contain diodes and to look at some simple and useful diode circuits. For a more
detailed description of semiconductor diodes and how they work, or if more precise
circuit analysis is desired, the reader should refer to a more advanced text.

6.1 Semiconductor Diodes

To understand how semiconductor diodes work, it is important to remember that an


electron cannot share its state with another electron. Hence, when there are a number
of possible electron states to fill, only one electron can go into each state. The lowest
energy state for an N electron system, the so-called ground state, will have the lowest
N energy states filled and the remainder of the states, all of which would have higher
energy, empty.
An atom will have discrete energy levels. When two such atoms are brought
together those energy levels split into two due to the interactions between the elec-
trons on each atom. A classical analog is the coupling of two harmonic oscillators
or pendulums. With a weak coupling (weak compared to the other forces in the
problem), what was a single resonant frequency for the oscillators becomes two
distinct frequencies. The process continues as more atoms (oscillators) are brought
together. Thus, if, say, 1023 atoms are brought near each other, such as in a solid, then
the levels break up into 1023 sublevels. This large number of levels, that are almost
of equal energy, are usually described using “band theory.” That is, rather than talk
about discrete levels as in an atom, one talks of a band of levels. While a discrete
level is either occupied or empty, a band can be fully occupied, completely empty,
or partially filled. The many states that make up a band that is in its lowest energy
configuration will still fill from lowest energy to highest.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 115
B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_6
116 6 Diodes

Fig. 6.1 Representation of


the uppermost filled and
lowermost empty energy
bands in a pure
semiconductor at absolute
zero

The energy bands for a pure (“intrinsic”) semiconducting material at absolute


zero (see Fig. 6.1) will have an uppermost energy band that is completely filled,
known as the valence band. Above the valence band is a lowermost band that is
completely empty, called the conduction band. In between is a region of energy
where there are no allowed states—that is, there is an “energy gap” between the two
bands. Completely filled or completely empty bands are “inert” in that they do not
contribute to electrical conduction at normal operating voltages—in much the same
way that filled electron shells make the noble gases inert to chemical reactions. In
the case of an empty band, there are no charge carriers to carry a current. In the case
of a full band there will always be as many electrons going in any given direction as
are going in the opposite direction, and hence no net current.
A semiconducting material is one where the band gap is “not too large.” When
the gap is large, the material is an insulator. There is no sharp cut-off between
semiconductors and insulators, but a practical value is that a band gap of ~0.3 to
~1 eV is desirable for a useful electronic device.1
At finite temperatures (such as room temperature) some electrons from the valance
band will be excited into the conduction band leaving behind an empty state, a “hole,”
in the valence band. The “hole” is the absence of an electron. The empty state
can move around and, compared to the inert state (completely full), it is positively
charged.
When referring to “electrons” in a semiconductor what is usually meant is “elec-
trons in the conduction band” only. All the other electrons are inert and do not be
considered. Likewise, a “hole” refers only to empty states in the valence band. If an

1 The electron-volt, or eV, is a non-SI unit of energy equal to the amount of energy gained by an
electron when accelerated through 1 V. 1 eV = 1.60218 × 10−19 J.
6.1 Semiconductor Diodes 117

electron and hole “recombine,” that is, an electron in the conduction band falls back
to fill a hole in the valence band, one is left with the inert state, or “nothing.” It is as
if both the electron and hole disappeared.
Electrical current in a semiconductor can be carried by electrons in the conduction
band and/or holes in the valence band. These charge carriers are free to move under
the influence of an applied electric field—there are plenty of states available for them
to occupy.
As a crude analogy, imagine a room filled with N chairs representing N possible
states. If N people occupy the room and all are sitting, there will be no empty chairs.
If the people are only allowed to change chairs one at a time, no one will be able to
move in such a room. However, if the number of people is small compared to N, then
they are pretty much free to move about. If it gets hot in the room, they all can move
toward the window, and so on. On the other hand, if there are just a few less than N
people in the room, there will be only a small number of empty chairs. The people
next to those empty spots can choose to move into the empty spot, causing the empty
spot to move in the opposite direction. If the room gets warm and the people try to
move to the window, the empty spot will move away from the window. That empty
spot is analogous to the hole in the semiconductor, but rather than talk about what
fills the hole, it is easier to talk about the hole, in this case the absence of a person,
as if it were the object of interest.
The conducting properties of the intrinsic (“i”) semiconducting material can be
modified by adding impurities. Adding one type of impurity effectively adds electrons
to the system and is called n-type doping (“n” for the negative charge). Adding another
type effectively removes electrons, the same as “adding holes,” and is called p-type
doping (“p” for the positive charge of a hole). The product of the electron density
and the hole density is (roughly) constant, so with strong n-doping there will be very
few holes around, and with p-doping there will be very few electrons available to
carry current.
No net charge is added to the system with impurity doping. The compensating
charge is contained in the immobile nucleus of the impurity atoms. That is, for n-
doping, there will be extra protons present and with p-doping, a deficiency of protons
when compared to the pure system (the “inert” state).
The basic semiconductor diode is the “p–n junction diode” and consists of a layer
of p-doped semiconductor immediately adjacent to a layer of n-doped semiconductor
(Fig. 6.2). The diode conducts from p to n relatively well, but does not conduct well
in the reverse direction.
The electrons and holes will be moving around at speeds near the speed of light
and can diffuse across the barrier. Of course, the “n” carriers will leave their extra
proton behind and the “p” carriers their missing proton—a negative charge compared
to the inert state. Hence there are forces trying to pull the electrons and holes back
into their own region. An equilibrium is established between the tendency to diffuse
and this electric field that pulls them back.
Near the p–n junction is a “depletion region” where electrons have diffused across
into the p-region and holes have diffused across to the n-region resulting in recom-
bination, i.e., “nothing.” The size of the depletion region varies considerably with
118 6 Diodes

Fig. 6.2 A semiconductor diode is constructed from layers that are “doped” with impurities created
an excess of holes (“p”) in the valence band and/or an excess number of electronics (“n”) in the
conduction band. In equilibrium, at the boundary there will be a depletion region with no mobile
charge carriers

conditions but is of the order of 1 µm across. Associated with the depletion region
is an energy barrier to cross the junction. That barrier arises from the equilibrium
electric field within the material. With no external electric field applied, the equilib-
rium is established so that the barrier is just barely high enough so that there is no net
current. When an electric field is applied in one direction, the barrier is lowered and
conduction occurs, when applied in the other direction the barrier is raised further
and, of course, no conduction occurs since the barrier was already high enough to
stop conduction.
The schematic symbol for a diode is shown in Fig. 6.3. The voltage across the
diode is taken to be positive if the voltage at the anode is larger than the voltage at
the cathode. The current through the diode is taken to be positive when it is in the
forward direction. When the voltage and current are positive, the diode is said to be
“forward biased.” When the voltage is negative, trying to push the current backward
through the diode, the diode is said to be “reverse biased.”
To understand electronic circuits containing diodes, the voltage-current relation-
ship for the diode is all that really matters. Since that relationship is somewhat compli-
cated for real diodes, simpler models are often used to describe the behavior. Which
model is chosen will depend on the demands of the application being considered.

Fig. 6.3 The standard


schematic for a diode. If the
triangle is considered an
arrowhead, it points in the
forward direction, where
conduction is easy. In the
reverse direction, there is a
wall to block the current
6.2 Diode Models 119

6.2 Diode Models

As stated above, a diode is a two-terminal device that conducts easily in one direction
and poorly in the other. It is, therefore, a non-linear device. In general, the basic
theorems for linear devices (superposition, etc.) cannot be expected to work for non-
linear devices. Even simple equation solving methods can become quite cumbersome,
if not impossible, to use when there are non-linear elements present. However, there
are methods that can be used to get practical results.
Kirchhoff’s laws must be obeyed in all circuits, and at the same time the current–
voltage relationship for each device, the device rules, must be satisfied. To solve
circuits with diodes, a current–voltage relationship for diodes is required. The most
accurate “rule for the diode,” that is, the relationship between the voltage across and
current through any diode, is known only graphically based on measurements. In
order to handle many non-linear devices, simplified models are used that approximate
the real behavior of the devices. Those models may be based in theory or may simply
be an ad hoc description of observed behavior. Each of those models will have a
range of validity which must be verified. Determining which model to use depends
on the situation, the accuracy required for the result, and the computational resources
available.
A real semiconductor diode does not start conducting appreciably until a small
“turn-on voltage” has been reached in the forward direction. This is typically 0.3–
0.7 V. Beyond that there will be some resistance within the device. These behaviors
are often approximated using simple models. The simplest models for calculations
involving diodes are based on piece-wise linear models. For each linear region the
problem can be treated as a linear circuit. The solution for each linear piece is then
stitched together to find the total behavior.

6.2.1 Piece-Wise Linear Diode Models

Table 6.1 illustrates three of the simplest models for a diode, the piece-wise linear
models. These are constructed using several straight line segments.
For the piece-wise linear models the circuit can often be solved by assuming that
the solution lies on one of the linear pieces. The (now linear) circuit is solved based
on that assumption, and then checked to see if that assumption is borne out by the
result. If the solution matches the assumption, then the problem is solved. If the
solution does not match, another linear piece of the model must be used.
Alternatively, of course, a graphical solution can be used. In this case an equation
is found for the circuit in question that relates the voltage and current across the diode,
V d and I d . These equations are obtained using Kirchhoff’s laws (the general rules).
The equations are plotted alongside the diode characteristics (the rules for the device).
Where these two curves intersect is where both rules are satisfied simultaneously and
120 6 Diodes

Table 6.1 Piecewise linear models for diodes


Model Description Graph of I versus
V
“Ideal diode” (the simplest model) Either V = 0, I > 0 or I = 0, V < 0 Id

Vd

Ideal diode with “turn-on” voltage Either V = V 0 , I > 0 or I = 0, V < Id


(Improves on the ideal diode if voltages V 0 (V0 ≈ 0.5 V for typical
encountered are less than or comparable semiconductor diodes, though
to about 1 V) somewhat larger for some special Vd

diodes such as LEDs)

Ideal diode with turn-on voltage and Either I = 0, V < V 0 or I = (V − Id


some resistance (Improves on the above V 0 )/Rd if V > V 0 (Rd is an
by adding some diode resistance. Use if effective resistance for the diode.
diode resistance is not negligible Its value depends on the “typical Vd

compared to others in the circuit) current” encountered)

is the solution. Accurate numerical solutions obtained using a computer are often
equivalent numerical implementations of such graphical solutions.

6.2.2 An Analytic Model for the Semiconductor Diode

An approximate expression relating the voltage across, V d , and current, I d , through


a real semiconductor diode is

Id = I0 (exp(Vd /ηVT ) − 1)
T V
VT = = 0.026 V at 300K, (6.1)
11600 K
where V T is the “volt equivalent of temperature” (temperature, T, should be in
kelvin). The parameter η (eta) will depend on the particular semiconductor and
is approximately 1 for most silicon diodes.
The constant out front, I 0 , is the “reverse saturation current,” which is also some-
what temperature dependent. If the temperature does not stray too far from room
temperature, it is not necessary to worry about that dependence here. The reverse
saturation current is the magnitude of the current for the diode in the reverse direc-
tion. For typical modern silicon diodes, I 0 can be as small as a few picoamps (pico-
is 10−12 ) and rarely exceeds a nanoamp (nano- is 10−9 ). The reverse current for a real
diode is not zero, but it is usually very small compared to most currents of concern.
6.3 Solving Circuits with Diodes 121

The expression above is reasonably accurate provided the diode current and/or the
voltage are not too large in magnitude—that is, as long as one stays well within the
ratings of the device—and if operation is not too far from room temperature. Even
though the expression above is reasonably simple, when used to solve for currents
in a circuit it is likely that either graphical or numerical techniques will be required
to find solutions to the transcendental equation(s) that result. Equation (6.1) will be
useful later for analyzing an op-amp circuit that contains a diode. That equation also
indicates there is some temperature dependence which may need to be taken into
account and/or can be calibrated and used to make temperature measurements.

6.3 Solving Circuits with Diodes

The choice of solution method for a circuit with a diode depends on the desired
accurately of the results. For most cases that are likely to be encountered, one of
the simple piece-wise linear models is more than accurate enough. When the diode
is simply used as an on/off device, that is, as a “rectifier,” the ideal diode model is
usually quite sufficient, at least to understand what the circuit does.
Since the diode is not linear, some care is necessary when solving circuits.
However, if a piece-wise linear model is used and conditions are such that the diode
stays on one linear piece, the diode can be treated as a linear device. Sometimes it
is not obvious ahead of time if this will be true. In such cases, a solution can be
found by guessing. The guess is checked once the solution is found. If everything
works out, then that solution is (probably) ok. If not, the guess was wrong and a new
solution must be found starting with a different guess.
If more precision is required, in particular if the “turn-on” voltage of the diode
is comparable to the accuracy needed, then it may be necessary to use graphical or
numerical techniques.

6.3.1 The Ideal Diode

Many circuits can be treated using the ideal diode model. Those that require one of
the other piece-wise linear models can be solved in the same way, since the piece-
wise linear models are equivalent to an ideal diode with the addition of some linear
components (e.g., resistors and/or batteries).
Consider the simple circuit shown in Fig. 6.4a. Assume V 0 is positive and enough
larger than the diode’s turn-on voltage so that the turn-on voltage can be neglected.
Using the ideal diode model, the diode is either conducting as if it were a wire or it
is an open circuit. These two possibilities are shown in Fig. 6.4b, c. For circuit (b),
the resistors are in series so the current is easily found to be I = V 0 /(R1 + R2 ) in
the direction shown by the arrow. For circuit (c), the current will be zero since there
is no complete circuit. Now go back to check which of these is consistent with the
122 6 Diodes

Fig. 6.4 The simple circuit in a is analyzed using the ideal diode models in (b) and (c). In b the
diode is considered conducting and in c it is considered to be in reverse bias. Only one of (b) or
(c) can be correct

ideal diode. The solution for (b) clearly works as the current found is indeed in the
forward direction, as assumed. The circuit at (c) has V ' = 0, and V out = 0. Since it
was assumed that V 0 > 0, the diode must be forward biased, which is not consistent
with the open circuit behavior. Hence the solution is the value obtained from circuit
(b) and not (c). For circuit (b) the series resistors form a voltage divider, so V out =
V 0 R2 /(R1 + R2 ). In practice the measured value will be about 0.5 V less than this
due to the turn-on voltage of real semiconductor diodes.
Diodes are often found in circuits involving time-dependent signals. They may be
used as part of a signal processing application or in a power supply for rectification of
incoming AC power (e.g., from a wall outlet) to create DC power. The circuit analysis
is the same in either case. Some examples are presented next and are discussed using
the ideal diode model.

6.3.1.1 Half-Wave Rectifier

The circuit in Fig. 6.5a is called a “half-wave rectifier.” It is perhaps the simplest
way to convert signals that may be both positive and negative into a signal that is
only positive (or by turning the diode around, only negative). Using the ideal diode
model as above, it is clear that when V in > 0 the diode conducts and V 0 = V in . When
V in < 0, the diode blocks the current and so with no current across R, it must be that
V out = 0. Hence, V out as a function of time (solid) when V in is sinusoidal (dotted)
looks as shown in Fig. 6.6a.

(a) (b)

Fig. 6.5 a A simple half-wave rectifier, shown with a sinusoidal source, lets through only the
positive part of the sine wave. In b the output is filtered by a capacitor so it will be closer to being
constant in time
6.3 Solving Circuits with Diodes 123

(a) (b)

Fig. 6.6 The output of the circuits of Fig. 6.5a, b respectively. In (b) traces for three different
capacitance values are shown. The larger the capacitance, the less droop will occur between maxima

The output represents only the positive half cycle of a sinusoidal input so this
circuit has come to be referred to as a half-wave rectifier.
The output of this half-wave rectifier has a non-zero average, but the output still has
a significant time dependence. To smooth out the time dependence, a low-pass filter
can be used. Perhaps the simplest filter is to add a capacitor, C, as shown in Fig. 6.5b.
When V in > V out the capacitor charges, however when V in < V out , the capacitor can
only discharge through the resistor. If the RC time constant is long compared to
the period of the sinusoidal input, the output will smooth out and will appear as
illustrated in Fig. 6.6b for three different capacitance values. Of the values shown,
the upper curve corresponds to the largest capacitance and the lower to the smallest.
When the capacitance becomes very, very large, the output will be essentially flat.
The half-wave rectifier finds use for more than just sinusoidal inputs. Such a
circuit can be used to extract the maximum amplitude of any time-dependent signal.
Another example is for use as protection against an externally supplied signal or
power source. If the source is negative when it should be positive, this circuit blocks
the negative signal from the remaining circuitry, thus preventing possible damage.

6.3.1.2 Diode Limiter

A limiter is useful in order to protect later circuitry from signals that are too large or
too small. A simple diode limiter is shown in Fig. 6.7a, where V max > V min . When
V in > V max the upper diode is forward biased and the lower reverse biased. The ideal
diode model would then predict behavior as shown in Fig. 6.7b, and hence V out =
V max . On the other hand, if V in < V min the behavior will be as shown in Fig. 6.7c and
V out = V min . When V min < V in < V max both diodes are “off” and so, using the ideal
diode model, the diodes each look like an open circuit. In that case V out = V in .
An example of how this circuit functions for a time-dependent signal is illustrated
in Fig. 6.8, where a hypothetical input signal, V in , is shown with the dotted curve and
the corresponding output voltage, V out , is the solid curve. When a signal is simply
truncated in this manner it is said to have been subjected to “hard limiting.” Imagine
an object moving in a room between a hard ceiling and a hard floor. “Soft limiting”
124 6 Diodes

(a) (b) (c)

Fig. 6.7 A limiting circuit shown in a is analyzed using the ideal diode model in b and c. If the
input exceeds V max or gets smaller than V min , the corresponding diode conducts

Fig. 6.8 An example


showing the effects of a
limiting circuit on a
time-dependent signal. The
original signal is shown
dotted, while the output of
the limiter is solid

would be more like an object moving vertically2 in a room where pillows were spread
across the floor and fastened to the ceiling.

6.3.1.3 Diode Clamp and Voltage Doubler

The circuit in Fig. 6.9a is referred to as a diode clamp. Using the ideal diode model,
when the input is negative the equivalent circuit is as shown in Fig. 6.9b. The capacitor
is charged to match the (negative) input voltage. When the voltage returns to positive,
the capacitor cannot discharge through the diode. If the RC time constant is long
compared to the period of the (sinusoidal) input, the capacitor will have comparatively
very little time to discharge at all. Since the capacitor will ultimately charge to
match the maximum negative voltage, when the input returns positive, that maximum
negative value is added (positively) to the input. The net result is that the signal is
translated so that what was the maximum negative voltage is now at zero. This is
illustrated in Fig. 6.10 for a sinusoidal input, where the input is shown dotted and
the output is the solid line.
If the amplitude of the input signal is changed, the minimum of the signal will
adjust so that it is always at zero—it is “clamped” at zero. The maximum of the

2 Imagine a child jumping on their bed.


6.3 Solving Circuits with Diodes 125

(a) (b) (c)

Fig. 6.9 The diode clamp circuit in a is analyzed in b and c for a negative input followed by a
positive input respectively

Fig. 6.10 The output from


the diode clamp of Fig. 6.9a
(solid) compared to the input
(dotted) for a sinusoidal
source

signal is then the peak-to-peak amplitude of the input. If the diode is reversed, the
signal is clamped so that the maximum is always zero. If desired, a constant voltage
source (e.g., a battery) can be added to the circuit to clamp the voltage at a non-zero
value.
If the clamp circuit is followed by a half-wave rectifier with a capacitor, the result
is similar to the filtered result found above for the half-wave rectifier by itself, but
with the peak amplitude now doubled. Such a circuit is called a voltage doubler and
would look like Fig. 6.11a. Additional diodes and capacitors can extend this idea to
create extremely large (DC) output voltages.3 Each stage of such a circuit consists
of two diodes and two capacitors. The output voltage across the resistor, R, for an N
stage multiplier, provided the capacitors are large enough, will be NV pp , where V pp
is the peak-to-peak voltage of the input signal. For example, a 3-stage multiplier is
shown in Fig. 6.11b. Such a circuit is sometimes referred to as a “voltage ladder.”

6.3.1.4 Full-Wave Rectifier

The half-wave rectifier can be used to turn AC power into DC power; however, half
of the available signal is unused. The full-wave rectifier, or absolute value circuit,

3Such a scheme was used in the 1930s by J. D. Crocket and E. T. S. Walton for their Nobel Prize
winning particle accelerator, and the so-called Crocket-Walton multiplier circuit shown here is
named for them.
126 6 Diodes

Fig. 6.11 A simple voltage doubler is shown in (a). The idea can be extended to achieve very high
voltages, such as shown in (b)

can be used instead. A simple full-wave rectifier based on a diode bridge is shown
in Fig. 6.12a.
As long as the magnitude of V in is significantly larger than the diode turn-on
voltage, the behavior can be analyzed using the ideal diode model. When V in > 0
the circuit behaves as shown in Fig. 6.12b, and when V in < 0 the circuit behaves as
shown in Fig. 6.12c. In each case, consider which way the input source is trying to
push the current. If it is in the forward direction, the diode is replaced with a wire, if
in the reverse, an open circuit. The result is that in all cases, the positive side of the
input is connected to just one side of the resistor, the negative side to the other, and
hence the output is always positive. For a sinusoidal input, the output as a function of
time will look like what is shown in Fig. 6.13, where once again the input is shown
dotted.
As was done for the half-wave rectifier, the output of the full-wave circuit can be
smoothed with the addition of a capacitor (Fig. 6.14a). In practice, it is important to
note that for the circuit shown, the input source and the output cannot both have a
ground connection. Remember that all grounds are connected to each other, even if

Fig. 6.12 The diode-bridge full-wave rectifier circuit shown at (a) is analyzed using the ideal diode
model in (b) and (c)
6.3 Solving Circuits with Diodes 127

Fig. 6.13 The output from the circuit of Fig. 6.12a (solid) compared to the input (dotted)

Fig. 6.14 a A capacitor is added to the full-wave rectifier to smooth the output. A common imple-
mentation uses a transformer on the input, as shown in (b), that can scale the voltage and eliminate
potential problems with the ground connections

that connection is not visible in the diagram. In power supply applications, the input
power may come from a wall outlet at a voltage level different from what is desired.
Both the grounding and voltage level can be addressed using a transformer on the
input (Fig. 6.14b). The transformer only passes the time-dependent signals and hence
the time independent ground level is blocked. As discussed earlier, a transformer can
also be used to change the amplitude of a time-dependent input signal. Such a circuit
is common for some types of inexpensive DC power supplies or as the first stage of
a more elaborate supply.
The diode bridge—four diodes wired appropriately for a full-wave rectifier—is
available as a single device. Those devices designed for higher current applications
include provisions that can be used to help dissipate the heat generated by real diodes.

6.3.2 Graphical Solutions

An important tool for understanding the behavior of all non-linear circuit elements
is the use of graphical techniques. Remember that all circuits must obey Kirchhoff’s
laws and each element of the circuit must obey its own rule. The strategy is then to
write down the conditions for each of those separately, graphing the two results on
a common grid, and then finding the spot where both rules are satisfied.
128 6 Diodes

Fig. 6.15 A simple example


circuit used as an example to
illustrate graphical solutions

Consider the circuit of Fig. 6.15 where an ideal voltmeter is used to measure V out .
First, write down the equations from Kirchhoff’s laws,

I1 = I2 + I3
V1 − I1 R1 − I2 R2 = 0 (6.2)
I2 R2 − Vd − I3 R3 = 0.

Note that the (as yet unknown) voltage across the diode is simply entered as V d . The
voltage across the diode is a function of the current through the diode, which in this
case is I d = I 3 . There is also a relationship between V d and I d , which is the “rule for
the diode.” To get the results of Kirchhoff’s laws into a form so they may be graphed
on the same grid, they need to be put into a form where V d is a function of I 3 = I d .
For this circuit, one way to do this is to replace I 1 in the second equation above
using the first equation. Now multiply the second equation by R2 /(R1 + R2 ),

R2 R2
[V1 − (I2 + I3 )R1 − I2 R2 ] = [V1 − I3 R1 ] − I2 R2 = 0, (6.3)
R1 + R2 R1 + R2

and add this to the third equation to eliminate I 2

R2
V1 − I3 (R1 ||R2 + R3 ) = Vd . (6.4)
R1 + R2

The result is a straight-line relationship between V d and I 3 = I d that must be obeyed


for this circuit in order to satisfy Kirchhoff’s laws.
Now plot the straight-line result from Kirchhoff’s laws on a graph of I d (= I 3 ) as
a function of V d . That line can be plotted by finding any two points on that line and
connecting them. It is perhaps most convenient to use the intercepts with the axes.
That is, put in V d = 0 and solve for I 3 to get one point, then put in I 3 = 0 and solve
for V d to get another point. Now connect those points with a straight line. Put the
device characteristics on the same graph. The point where the two curves intersect
is the solution—both the general and device specific rules are satisfied there.
6.4 Diode Ratings 129

Fig. 6.16 To solve the


circuit of Fig. 6.15, the diode
characteristic (solid) and the
results obtained from
Kirchhoff’s laws (dashed)
are plotted together. Since
both must be satisfied
simultaneously, the only
solution is at the intersection
of these curves

6.3.2.1 Example Graphical Solution

Consider a diode curve obtained from direct measurement of a real diode (solid curve
in Fig. 6.16), and use the circuit above as an example, with V 1 = 3 V, and R1 = R2
= R3 = 1 k. Then the straight line from Kirchhoff’s laws is

Vd = 1.5V − 1.5k · I3 , (6.5)

so if V d = 0, I 3 is 1 mA, and if I 3 = 0, V d = 1.5 V. The resulting line is shown


dashed in Fig. 6.16.
The only spot that satisfies both the overall rules and the device rule will be the
solution. The solution here is then I d = I 3 = 0.6 mA, V d = 0.6 V. These values are
substituted into the Kirchhoff’s law equations to get I 1 = 1.8 mA and I 2 = 1.2 mA.
This type of graphical method can be used with any two-terminal device, no matter
how strange, if measured device characteristics are available.

6.4 Diode Ratings

The two diode ratings that are often important are the “peak inverse voltage (PIV)”
and “maximum forward current.” The PIV is the largest voltage one can expect to use
in the reverse direction before breakdown occurs. The maximum forward current is
the largest current one can expect to get through the diode before damaging the device
(principally due to the heat generated). Sometimes a maximum forward voltage is
also specified.
130 6 Diodes

6.5 Diode Capacitance and Response Time

Since a semiconductor diode has adjacent layers with opposite charges (electrons
and holes) separated by a non-conducting region (the depletion region), a diode can
have a significant capacitance (a few pF). This capacitance, along with any resistance
in the circuit that contains the diode, form an RC time constant. This RC behavior
is one important factor that limits how quickly the diode can respond to a change in
the input conditions. If speed is of concern, some capacitance may need to be added
to the diode model to get accurate predictions from the circuit analysis.

6.6 Specialty and Other Diodes

There are various other diodes that have a more specialized purpose. A summary of
many of them is included in Table 6.2. While specialized schematics are often seen
for these diodes, the simple diode schematic (Fig. 6.3) is sometimes used as well.
Two of these will be considered in more detail below: Zener diodes and light-emitting
diodes (LEDs).

Table 6.2 Some specialty diodes


Schematic Name Brief description
Zener diode Designed to be used at the reverse breakdown voltage as a
limiter or voltage reference. Reverse break-down current is
limited by power considerations
Light emitting A semiconductor diode optimized to emit light due to
diode (LED) electron–hole recombination when the diode is conducting in
the forward direction. Electrically similar to other
semiconductor diodes, but usually with a larger turn-on
voltage (Can be used in reverse as a light detector)
Photodiode A semiconductor diode optimized to detect incident light. The
diode is used in series with a resistor and with reverse bias.
Incident light creates electron–hole pairs and increases the
conductivity of the diode
Varactor diode Designed to use the inherent capacitance of a semiconductor
diode due to the depletion region. By varying a DC biasing
voltage in the reverse direction, the size of the depletion
region, and hence the capacitance, can be adjusted
Schottky diode Generally has a smaller turn-on voltage and a very rapid
turn-on time compared to other semiconductor diodes. Used
when either of those parameters is a significant issue. Relies on
a metal–semiconductor junction. Also known as a barrier
diode or a hot-carrier diode. Not to be confused with the
four-layer Shockley diode
(continued)
6.6 Specialty and Other Diodes 131

Table 6.2 (continued)


Schematic Name Brief description
Tunnel diode Based on quantum mechanical tunneling, and has the unusual
property that for some operating conditions the dynamic
resistance is negative—that is, the current decreases when the
voltage across it increases
Laser diode As its name implies, a laser diode is constructed to produce
laser light. Usually based on a p-i-n layer structure. The laser
light arises from the area near the junction. Often represented
in a circuit using the simple diode schematic, without the arrow
Vacuum tube A diode based on vacuum tube technology. Largely obsolete
diode except for novelty use and for very high-voltage and/or very
high-power devices. Otherwise, used like other diodes

All diodes have a reverse breakdown voltage. If the reverse voltage gets too large
in magnitude, the diode begins to conduct. In many cases, this leads to destruction of
the device. Zener diodes are designed to do this in a controlled way so that, in a sense,
they have a very large turn-on voltage in the reverse direction. The reverse conduction
can occur through a quantum tunneling process (the Zener effect) or through an
avalanche process. The avalanche starts with one electron (or hole) breaking free,
and that causes more to break free and so on. The reverse voltage where reverse
conduction occurs is controlled during the manufacture. Whether the Zener effect or
the avalanche dominates, the diodes are usually referred to as Zener diodes4 and the
voltage where it occurs is referred to as the Zener voltage of the device.
Zener diodes are specified by the (magnitude of the) reverse voltage where conduc-
tion begins. Examples of measured values for several Zener diodes are shown in
Fig. 6.17. In the forward direction, they behave the same as a normal semicon-
ductor diode. In the reverse direction they behave normally until the Zener voltage is
reached.5 The almost vertical behavior is then well-modeled with a constant voltage
source (a battery).
Zener diodes are used for a number of purposes. Two Zener diodes in series, but
pointed in opposite directions, can be used to make a simple limiter. In addition,
Zener diodes are often used as voltage references and, for lower power circuits, these
diodes can be used as a simple voltage regulator. Note that the power dissipated in
the diode is the product of the voltage and the current. Since the voltage can be large,
the current is generally kept small.
When electrons and holes recombine, energy equal to the energy gap must go
somewhere. The two likely places are into a phonon (a quantized vibrational state
of the atoms, and ultimately heat) or into a photon of light. A diode optimized to
produce light, encased in a transparent case, can be used as a light source. The color

4 References to an “avalanche diode” may still be found from time to time.


5 For most such diodes, the avalanche effect tends to dominate at higher Zener voltages, whereas
the Zener effect dominates at lower Zener voltages. The Zener effect has a more gradual “turn-on”
characteristic.
132 6 Diodes

Fig. 6.17 The measured


(DC) characteristic curves
for 3.3 V (dashed), 5.1 V
(dotted), 7.5 V (solid), and
12 V (dot-dash) zener diodes

Fig. 6.18 The measured


characteristic curves for
several small LEDs which
produce different
wavelengths of light

of the light depends on the size of the band gap.6 Such diodes are referred to as
LEDs, which is short for light emitting diodes. Within the normal operating range,
the power of the light emitted (or the number of photons per second) goes like the
current through the device. Small signal LEDs are used with currents of 1–20 mA.
Those used for higher power applications, such as home lighting and automobile
brake lights, require much more current.
Electrically, an LED looks like a regular semiconductor diode, except with a
larger turn-on voltage. Example characteristics are shown in Fig. 6.18 for several
small LEDs.

6 There are exceptions to this. For example, white LEDs may be created using fluorescent materials
that change the color and so the observed light does not come directly from the electron–hole
annihilation.
6.7 Problems 133

If an LED is reverse biased, light can enter and, through the reverse process, create
an electron–hole pair. That pair can then carry current. The reverse current is then
a measure of the incident light intensity. Diodes optimized for such detection are
referred to as “photodiodes.” When interaction with light is undesirable, the casing
material should shield the diode from light.

6.7 Problems

1. A full-wave rectifier (Fig. 6.P1) is constructed for use as a power supply. Use the
ideal diode model to predict the output which will be seen as a function of time
on an oscilloscope, for each of the following individual faults. Assume the power
from the wall is sinusoidal and when working correctly, the RC time constant at
the output is short compared to the period of the sinusoidal input.
(a) The transformer is not plugged in.
(b) Diode D1 has failed and acts like a short circuit.
(c) Diode D1 has failed and acts like an open circuit.
(d) Diode D2 has failed and acts like a short circuit.
(e) Diode D2 has failed and acts like an open circuit.
(f) Diode D2 was placed in the circuit backward.
(g) All the diodes were placed in the circuit backward.
(h) The resistor has burned out and is an open circuit.

Fig. 6.P1 Problem 1

2. A mystery device, “X,” (Fig. 6.P2) that has the measured voltage-current rela-
tionship shown in Fig. 6.P2c, is placed in the circuit shown. What is the current,
I, in the circuit?
3. Use the analytic model for the diode to predict the voltage across a silicon diode
when the current is 0.01, 0.1, and 1.0 mA. Assume room temperature, I 0 = 1
pA, and η = 1.
4. A 7.5 V Zener diode with a maximum power rating of 1 W is to be used in a
circuit. What is the maximum (sustained) current that can go through the diode?
5. For the circuit of Fig. 6.P5, that contains three identical small LEDs, describe the
behavior before and after the switch is closed. See Fig. 6.18 for representative
current–voltage relationships for LEDs.
134 6 Diodes

Fig. 6.P2 Problem 2

Fig. 6.P5 Problem 5

6. (Challenge Problem) Consider the circuit in Fig. 6.P6, where the capacitor is
initially uncharged and the switch is closed at t = 0. Assuming ideal components,
show that if L is large enough, the capacitor will become charged to close to
2V 0 with virtually no loss of energy in the resistor. In the process, you need to
determine and state clearly what is “large enough.”

Fig. 6.P6 Problem 6

7. (Challenge Problem) How might you measure the reverse current vs. voltage
characteristics for a diode using a real ammeter and voltmeter, given that the
reverse current is so small?
Chapter 7
FETs

The circuit elements considered thus far were two-lead devices. When describing the
behavior for those devices the relationship between just one current and one voltage
was required. Here the discussion moves on to transistors that are three-lead, non-
linear devices. Field effect transistors will be considered in this chapter and bipolar
junction transistors will be treated in the next. These two different types of transistors
serve as examples as to how more complicated devices can be treated using linearized
models.

7.1 Junction Field Effect Transistors

There are several different types of field effect transistors (FET) in use. When one
simply refers to a(n) “FET,” it is likely that one is referring to one based on a simple
p–n junction, but with three wires instead of 2, as shown in Fig. 7.1a. A FET made
with a simple junction is sometimes called a JFET (“J” is for “Junction”).
FET is either pronounced as three separate letters (F-E-T) or as a word that rhymes
with “bet” and “set.” When a prefix is added to describe what kind of FET it is, it is
almost always pronounced as a two-syllable word. That is, JFET would be “jay-fet.”
For a JFET, the side of the junction with one lead is called the “gate.” The other
side is the “channel” and the channel has two leads—the “drain” and the “source.”
Conduction through the channel is controlled by the gate. In this simple picture the
drain and source look interchangeable, though in practice a real device works much
better one way than the other.
The device shown is an “n-channel JFET.” The schematic symbol for the n–
channel JFET is shown in Fig. 7.1b. Swapping the p and n regions will make a
p-channel FET. The schematic for the p-channel JFET looks the same as for the
n-channel except the arrow points in the opposite direction. The arrow is on the gate
and points from “p toward n.” If the arrow is drawn off center, it is closer to the
source. Most of the examples used here are based on the more common n-channel
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 135
B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_7
136 7 FETs

Fig. 7.1 a Basic


construction of an n-channel
junction field effect
transistor (JFET) and b the
corresponding schematic
symbol

FET. For the characteristics of a p-channel FET, add a minus sign to all FET voltages
and currents.
In normal JFET operation, the p–n junction is reverse biased. That means the
current into/out of the gate will be very small (nano- or picoamps). If forward bias
is applied to the p–n junction, the gate current becomes large and the device may be
damaged.
With any three-lead device there are three electric potential differences (voltages)
and three currents that can be specified. In this case the voltages are the drain to gate
voltage (V dg ), the drain to source voltage (V ds ) and the gate to source voltage (V gs ).
The three currents are the currents into each lead, I d , I g , and I s . By convention, these
currents are taken as positive for currents going into the device.
For FETs under normal operating conditions, the gate current is often negligible,
so as an approximation, I g = 0. This means that, using Kirchhoff’s current law, I s
= –I d . That is, in most circumstances, only one current need be specified. Also, it
must be that V dg + V gs = V ds so only two voltages need to be specified. For a FET
the two voltages V ds and V gs are usually the two specified. Under normal operating
conditions an n-channel JFET will have Vgs ≤ 0 and Vds > 0.
A JFET works because of the changes in the size of the depletion region as a
larger and larger reverse bias is applied to the p–n junction, as illustrated in Fig. 7.2.
If a large enough negative bias is applied, the transistor will “pinch-off” and no
current can flow. That value is known as the pinch-off voltage, V p . When the applied
bias is smaller than this, there will be current through the channel. If the current is
too large, the voltage drop in the channel will result in pinch off—i.e., no current.
Hence the current is self-restricting and cannot get too large. This internal feedback
mechanism limits the current to an equilibrium value that is almost independent of
the drain to source voltage. For a more detailed and more precise description of the
physics behind JFET operation, refer to a more advanced text.
With the simplification that the gate current can be neglected, the FET “device
rules” require the relationship between three values—two voltages and a current.
Such a relationship can be described using a two-dimensional plot with a set of
curves. The two common ways of presenting the FET characteristics are to plot I d
versus V ds for different values of V gs or to plot I d versus V gs for different values of
V ds . The latter is known as the “transfer characteristic.”
A typical device characteristic for a low-power n-channel FET looks like the plot
in Fig. 7.3. Note that for V gs > 0 the gate is forward biased and those values are not
7.1 Junction Field Effect Transistors 137

Fig. 7.2 Illustrating the effects on the depletion region as the reverse bias is increased from a to c

Fig. 7.3 Characteristic


curves for a particular
n-channel JFET

used. By extrapolation, it can be seen that for this transistor the pinch-off voltage,
the value of V gs where the drain current, I d , goes to zero, is about –1.8 V.
The so-called “active region” for this transistor, where the curves are roughly
horizontal, corresponds to (roughly) V ds > 2 V. For a JFET it is usually the case that
the active region corresponds to, very roughly, |V ds | > |V p |. Note that for an n-channel
JFET in the active region, both V gs and V p will be negative; for a p-channel JFET,
they are both positive.
An approximate formula, a mathematical model, can be used for the JFET in the
active region away from the origin. It is the parabola,
( )
Vgs 2
Id = Idss 1 − , (7.1)
Vp

where I dss and V p are device parameters. The approximation is valid only for 0 ≤
|V gs | ≤ |V p |, and where |V ds | is large enough to be in the active region. Do not try to
use this formula for other types of FETs without verifying its validity.
The value of I dss is the drain current when V gs = 0, which is the maximum current
one can expect. For the transistor characteristic shown in Fig. 7.3, I dss ≈ 5.4 mA.
138 7 FETs

The value of V p is the voltage when I d first goes to 0. As mentioned above, for the
curves shown in Fig. 7.3, V p ≈ –1.8 V. The values for I dss and V p depend somewhat
on the particular operating conditions, and “typical values” are usually quoted. Due
to device-to-device variations, those typical values are as accurate as one can expect
to have in any case.

7.2 Circuit Analysis with a JFET

There are two types of analysis for FET (and other transistor) circuits: a large signal
analysis, that is necessary when the non-linearities of the device are significant, and
small signal analysis based on a linear model for small changes from a particular
known starting solution. For a typical transistor amplifier both types of analysis are
necessary. The first establishes an “operating point” and the second deals with small
changes from that operating point. This same strategy is used in many areas of math
and science—solve most of the problem one way then treat the rest as a small change,
or perturbation.
The following two examples illustrate the analysis for larger signals. The first
illustrates a design problem, where the components to achieve a given outcome are
determined. The second illustrates an analysis problem for an existing circuit. That
is, given a set of components, what is the outcome?

7.2.1 Example 1—Determine Circuit Components

Consider the circuit in Fig. 7.4, where the JFET characteristics are those of Fig. 7.3.
If V = 15 V and it is desired to have V gs = –1 V and V ds = 9 V, what resistor values
should be used?

Fig. 7.4 Circuit for


Examples 1 and 2
7.2 Circuit Analysis with a JFET 139

Reading off the graph for the device characteristics in Fig. 7.3, if V ds = 9 V and
V gs = − 1 V, then I d = 1.6 mA.
The gate current will be very small so the value of Rg is not too critical. There
needs to be some path for that very small current, however. A value of Rg = 1 My
is usually more than sufficient to provide that path. Since the reverse current will
be nano- or picoamps, the voltage drop across 1 My is still small enough to be
negligible. Thus, the voltage at the gate will be the same as the ground connection,
or V g = 0.
The voltage at the source, V s , compared to ground, can be found from Ohm’s law.
The voltage drop from the source to ground is I d Rs . Hence,

Vgs = Vg − Vs = 0V − 1.6mA · Rs , (7.2)

and so if V gs = –1 V, Rs = 1 V/1.6 mA = 625 y.


Using Kirchhoff’s voltage law (KVL) from the power supply to ground,

15V − Id Rd − Vds − Id Rs = 0, (7.3)

and so

Rd = (15V − 9V − 1V)/1.6mA = 3100y. (7.4)

If one or both of the computed resistor values had come out negative, then it would
not have been possible to get the desired values using this circuit.

7.2.2 Example 2—Determining the Operating Point

Consider the same circuit with V = 15 V, Rd = 2400 y, Rs = 400 y, and Rg = 1


My. What are V gs , V ds , and I d ?
Once again, the current into the gate is negligible so V g = 0 V.
There are two different approaches that can be used at this point: a graphical
solution or the use of the JFET mathematical model. Since the graphical solution
technique is more general and will always work, that will be considered first.
Kirchhoff’s voltage law on the drain-source side of the transistor gives

15V − Id Rd − Vds − Id Rs = 0, (7.5)

or

Id = (15V − Vds )/(Rd − Rs ) = 5.4mA − Vds /2800y . (7.6)


140 7 FETs

Fig. 7.5 To solve Example 2 graphically, the characteristic curves and load line in (a) are used to
construct the transfer characteristic in (b). The bias line, from Kirchhoff’s laws is added to (b) and
the intersection is the solution

This is a straight-line relationship between I d and V ds and is known as the “load line.”
That line can be plotted along with the device characteristics. See Fig. 7.5a. Whatever
the solution, it must be somewhere along that line. The intersections between the
load line and the device characteristic curves, shown with solid circles, can be used
to create a transfer characteristic. Those points are put on a plot of I d versus V gs
and connected with a smooth curve as shown in Fig. 7.5b. The solution must be
somewhere along that curve.
Now look at Kirchhoff’s voltage law from the gate to source. This is called the
bias line,

0V − Vgs − Id Rs = 0 or Id = −Vgs /400y, (7.7)

which is a straight-line relationship between I d and V gs . Put that line on the second
plot—the dashed line shown in Fig. 7.5b. The solution must be along that line.
Both conditions are satisfied at the point where the solid curve and the dashed line
intersect, and so that point must correspond to the solution. In this case, it is close to
I d = 2.2 mA and V gs = –0.85 V. Using those values in the load line equation,

Vds = 15V − 2.2 mA · 2800 y = 8.8V . (7.8)

The process can be simplified a bit by assuming the solution is in the active region
and noting that the curves in that region are reasonably flat. Then, rather than using
the load line to extract data for the second plot, a simple vertical line will suffice. For
example, simply look at the points corresponding to a fixed value of V ds somewhere
in the middle of the active region, say V ds = 9 V. That is, the points in Fig. 7.5a that
are indicated with open circles can be used instead of the solid circles. As long as
the transistor is in the active region, the solution will be the same, at least within the
7.2 Circuit Analysis with a JFET 141

accuracy obtainable by this method. The assumption that the solution is in the active
region will need to be checked at the completion of the calculation.
A second solution method uses the JFET mathematical model. This method may
not work well for other types of transistors. The first step is to extract the two
parameters I dss and V p from the device characteristics, assuming the active region.
For the transistor here, this was already discussed above with the results that I dss =
5.4 mA and V p = –1.8 V.
The solution should then simultaneously satisfy Kirchhoff’s voltage law and the
model. That is,

Id = −Vgs /400y (7.9)

and
( ) ( )2
Vgs 2 Vgs
Id = Idss 1 − = 5.4 1 − mA. (7.10)
Vp −1.8V

Putting the first into the second,


( )
Id 0.40 ky 2
Id = 5.4 1 − mA (7.11)
1.8V

and so with I d in mA,

0.267Id2 − 3.40Id + 5.4 = 0. (7.12)

This is a quadratic that is solved using the quadratic equation. That is,

3.40 ± 3.402 − 4 × 0.267 × 5.4 3.40 ± 2.41
Id = = = 1.86 or 10.9 mA.
2 × 0.267 0.533
(7.13)

Only the smaller solution is valid. The larger solution is out of the range where the
model is valid and should be discarded.
The value obtained using the mathematical model and the value obtained directly
from the graphical data differ by about 15%. Such a disagreement should be expected.
The model is not perfect. The value from the data is likely more accurate and should
be used when possible.
142 7 FETs

7.3 The FET A.C. Model

A very simple model for the low frequency behavior of an n-channel FET transistor
operating in its active region about some fixed operating point is shown in Fig. 7.6,
where only linear circuit elements are used. Since there is only a negligible gate
current, the model shows the gate lead with no connection—thus there will not be
any gate current. The transistor includes a dependent current source where the current
is proportional to the gate to source voltage, with proportionality constant “gm .” A
model resistor, r d , is also included, though in many applications its value is large
enough that it can be omitted. A p-channel FET would look identical except the
current source would point the other direction.
Note that the model parameters (here gm and r d ) will depend on the operating point
used—that is, on the large signal analysis—and also that the model only applies to
the changes from the operating point, not to the total gate to source voltage, total
drain current, etc. That is, V gs in Fig. 7.6 is really only the change in V gs , which
might be written as /\V gs , however the “/\” is typically omitted.
This linear model only works well when the changes from the operating point are
small. The larger the input signals, the less accurate the model.
The “transconductance” (also called the “mutual conductance”) gm , is given by
|
d Id ||
gm = y f s = , (7.14)
d Vgs |Vds at op. pt.

and the model drain resistance is given by


|
d Vds ||
rd = 1/yos = . (7.15)
d Id |Vgs at op. pt.

That is, gm is a measure of how much the drain current changes for a given change of
V gs (e.g., as one moves vertically on the characteristic curves) and 1/r d is a measure

Fig. 7.6 A simple AC


model for the JFET used for
small changes from a fixed
operating point
7.3 The FET A.C. Model 143

of the slope of the characteristic curves in the vicinity of the operating point. The
alternate symbolic names shown (using y’s with subscripts) are also common.
Values for gm and r d can be determined graphically from the characteristic curves.
The value of gm (but not r d ) can also be determined from the model (7.1) relating I d
and V gs .

7.3.1 Example—Determine Transistor Parameters

Consider the characteristic curves in Fig. 7.7 where the operating point (the result
of the large signal analysis) has been determined (or chosen) to be at V gs = –1.5 V,
V ds = 8 V. The appropriate model parameters for that operating point are determined
using neighboring values read off the graph (also shown in Fig. 7.7):

7.7 − 3.5 mA
gm ≈ = 4.2 mmho = 4.2 mS
(−1) − (−2) V
(7.16)
12 − 4 V
rd ≈ = 16 ky .
5.4 − 4.9 mA

The units used for gm are those of electrical conductivity: 1 mho = 1/(1 y) = 1 S,
and 1 mmho = 1 milli-mho = 1 mS = 1/(1 ky). The “mho” (ohm spelled backward,
and pronounced like the name Moe) is an older unit for conductance (1/resistance)
and has been replaced by siemens.1 Note that the abbreviation for siemens is an
uppercase “S” and must not be confused with seconds, which uses a lowercase “s.”
Units (and prefixes) are case sensitive.
There may be considerable variability in the transistor parameters, as much as
a factor of two or more, from one transistor to the next even when they are of the
same type. The curves given by manufacturers are “representative.” Hence, there is
usually no point in trying to get more than two or, in the best cases, three decimal
places from any solution.
In Fig. 7.8, the simple linear model (dashed lines) for changes from the chosen
operating point, derived from the graphically determined parameters, are compared
with the actual transistor characteristics (solid lines). The model may be somewhat
crude, however the fact that it contains only linear circuit elements makes it very
useful. Note that the farther from the operating point, the poorer the agreement
between the model and the actual characteristics.
Remember that the model only describes changes from the operating point.

1 Named for the German engineer, Werner von Siemens (1816–1892). Note that his name includes
the final “s” so that the unit should also include the final “s.” That is, the singular is “1 S” and not
“1 siemen.” After all, 1 °C is not “one degree celciu.”
144 7 FETs

Fig. 7.7 The parameters for


the AC model of Fig. 7.6 can
be found graphically from
the device characteristic
curves

Fig. 7.8 A comparison of


the measured characteristic
curves (solid line) and the
approximate behavior of the
model of Fig. 7.6 using
parameters derived
graphically as in Fig. 7.7
(dashed line)

At higher frequencies, capacitance within the transistor becomes important and


will need to be included in the transistor model. Other additions to the model may
be necessary in some other circumstances.
7.4 FET Amplifier Configurations 145

7.4 FET Amplifier Configurations

Figure 7.9 shows the three common configurations used for single-transistor FET
circuits used to amplify time-dependent signals. They all are closely related, the only
difference being where the input, output, and ground are connected. The DC (large
signal) analysis, to find the operating point, for all of these is the same and is as was
described in the previous section. For DC signals, the capacitors are open circuits
and hence can be removed.
Once the DC analysis has been performed, the parameters gm and r d (if necessary)
are used to look at small time-dependent changes from that point. That is, a linear
FET model is used. This linear analysis works only if the transistor is in the active
region and the input signals are small enough.
For the sake of simplicity, assume that the capacitors are large enough so that, at
the frequencies of the signals to be amplified, they can be replaced by a short circuit
(a wire). This is certainly not always the case, but in order to keep the amplifier
analysis simple, that will be assumed here. Note that any constant (DC) voltage has
no time-dependent part and thus is equivalent to every other DC voltage that has no
time-dependent part, including ground (0 V). Hence for the AC (time dependent)
analysis all DC voltages look like ground. For the AC analysis the three amplifier
configurations will then reduce to the simplified circuits in Fig. 7.10.

Fig. 7.9 The three basic configurations for single transistor JFET amplifiers

Fig. 7.10 Models for the circuits of Fig. 7.9 for small changes from a fixed operating point
146 7 FETs

Table 7.1 Voltage gain and input/output impedances for the three JFET amplifier configurations
Configuration Voltage gain = V out /V in Input impedance Output impedance
Common source −gm (Rd ||rd ) Rg Rd
gm (Rs ||rd ) Rs ||rd
Common drain 1+gm (Rs ||rd ) ≈1 Rg 1+gm (Rs ||rd ) ≈ 1
gm
(source follower)
Rd
( )
Common gate Rs +(rd +Rd )/(1+gm rd ) Rs || rd +Rd
≈ Rs || g1m rd + (1 + gm rd )Rs
1+gm rd

It is then straightforward to compute the results shown in Table 7.1 from these
models, where the “≈” signs are valid for typical, but not all, applications.

7.4.1 Example—Gain for Common Drain Amplifier

This example shows how to compute the voltage gain for a common drain amplifier.
Consider the AC model for a common drain amplifier. The capacitors are assumed
large enough so they can be replaced with wires and it is assumed that the transistor
is in the active region where the simple FET model applies. In this case the desired
quantity is the voltage gain, Av = V out /V in . The circuit can be redrawn, as shown in
Fig. 7.11 to aid in the analysis.
Now

Vgs = Vg − Vs = Vin − Vout


Vout = Id (Rs ||rd ) = gm Vgs (Rs ||rd ) (7.17)
= gm (Vin − Vout )(Rd ||rd ),

so

Vout (1 + gm (Rs ||rd )) = gm Vin (Rs ||rd )


Vout gm (Rs ||rd ) (7.18)
AV = = .
Vin 1 + gm (Rs ||rd )

Fig. 7.11 The common drain circuit of Fig. 7.10 rearranged for analysis
7.4 FET Amplifier Configurations 147

7.4.2 Example—Impedances for Common Gate Amplifier

This example shows how to compute the input and output impedances for a common
gate amplifier.
Amplifiers will be connected to other circuitry at both the input and output. To
determine how the amplifier affects this other circuitry, the amplifiers can be replaced
by their Thevenin equivalent circuits as seen looking into either the input or output.
The small signal models are used for these calculations. The problem can become
a bit more complicated than previous Thevenin equivalent examples because the
current source within the model is a dependent source—it cannot simply be turned
off under all circumstances. The common gate amplifier is a good example where
care must be taken.
The common gate amplifier with an input source, modeled by its Thevenin equiva-
lent (source V 0 , impedance R0 ) and where the output is connected to a load, modeled
by a simple resistor, looks like the circuit in Fig. 7.12. The same assumptions used to
compute the amplifier voltage gain are used here. To compute the input and output
impedances, a test voltage, V t , is applied (on paper) to the input and output respec-
tively and the resulting current, I t , is computed. The ratio of the voltage to current is
the impedance.
The models for the two impedance calculations are shown in Fig. 7.13. In both
cases, the gate voltage is zero and, since it is not needed for the calculation, the gate
connection is not shown.
Applying straightforward linear analysis (KVL and KCL),
( )
rd + Rd ||R L
Z in = Rs || (7.19)
1 + gm rd

and

Z out = Rd ||[(1 + gm rd )(Rs ||R0 ) + rd ]. (7.20)

Fig. 7.12 The common gate amplifier with a source and load attached
148 7 FETs

Fig. 7.13 The common gate amplifier with the JFET model set up to determine a the input
impedance and b the output impedance of the amplifier

7.5 The Ohmic Region

The discussion above concentrated on the transistor in the active region. Sometimes
the ohmic region, at lower values of V ds , is of use. In that region, the transistor can
be used as a voltage variable resistor. As V gs is changed, the (average) slope of the
line changes, and hence the effective resistance changes. See Fig. 7.14. Of course,
the FET will act like a resistor only for smaller changes in V ds , where the line is
relatively straight, and care is required to avoid values that are negative.

7.6 MOSFETs

MOSFETs use a very thin metal oxide barrier between the gate and the rest of
transistor. This oxide layer is an insulator and hence the gate current is even smaller
than the gate current for a JFET. In addition to an extremely small gate current, the
gate voltage may be both positive and negative. In a MOSFET the conducting channel
is controlled by the electric field created by a gate to source voltage. “MOSFET” is

Fig. 7.14 A close-up of the


ohmic region of the
characteristic curves of
Fig. 7.3
7.6 MOSFETs 149

Fig. 7.15 Basic construction


of an enhancement mode
MOSFET

almost always pronounced as a two-syllable word (Moss–fet). MOSFETs are also


referred to as IGFETs—insulated-gate FETs. IGFET is a more general term.
There are two basic types of MOSFETs: the enhancement and the depletion
MOSFET.
In the enhancement MOSFET, impurities are introduced so that there is no channel
when the gate to source voltage is zero. When a voltage is applied, a channel is created
by the minority carriers. For example, a p-channel MOSFET will be made from an
n-doped substrate. When a negative gate to source voltage is applied, a net positive
carrier concentration is induced near the gate, thus creating a p-channel in the n-doped
material. This is illustrated in Fig. 7.15.
In the depletion mode MOSFET a conducting channel in the absence of a gate
voltage is created by impurities. The gate voltage then modifies the width of the
channel. Figure 7.16 illustrates an n-channel depletion mode MOSFET with both
a negative (a) and positive (b) gate voltage. When the negative voltage is applied,
the channel size is reduced, thus reducing the current. When a positive voltage is
applied, the size of the channel is increased, increasing the current.
In addition to low-power applications, MOSFETs can be made for high volt-
ages (hundreds of volts) and large currents (>100 A) and hence they are good for
high power applications. Enhancement mode devices are commonly used in on/off
switching applications. Depletion mode devices are nice for amplification of AC
signals since the gate to source voltage can be both positive and negative.
Schematic representations of some MOSFETs are shown in Fig. 7.17. The very
thin oxide layer of the MOSFET is easily damaged by static electricity and many
devices include a built-in protection diode to reduce the possibility of damage. That
protection diode may appear within the schematic for the MOSFET.

Fig. 7.16 The construction of a depletion mode MOSFET shown at a with V gs > 0 and in b with
V gs < 0
150 7 FETs

Fig. 7.17 Schematic representations of n- and p-channel MOSFETs. Depletion mode devices are
shown in a and enhancement mode in b with n-channel devices on the left and p-channel devices
on the right in both cases

Fig. 7.18 Measured device


characteristics for a
particular enhancement
mode MOSFET

Characteristic curves for two sample MOSFETs are shown in Fig. 7.18 (enhance-
ment mode) and Fig. 7.19 (depletion mode). These curves can be used to obtain
graphical solutions in a manner identical to the curves for the JFET. There is nothing
particularly special about these two examples—they were chosen to illustrate some
of the variety in behavior that can be expected from different transistors.

7.7 Additional Application—Dynamic Memory

A capacitor connected to a MOSFET gate can hold its charge for a relatively long
time. Figure 7.20 shows how this fact can be used to create a simple 2-state memory
element from an enhancement mode MOSFET.2 To store a value, the capacitor is
momentarily connected to either +V (to charge it) or ground (to discharge it). When
disconnected from both, the capacitor will hold its state of charge for a time governed

2 The configuration shown is for demonstration purposes only. The configuration commonly used
in large memory arrays differs from what is shown.
7.7 Additional Application—Dynamic Memory 151

Fig. 7.19 Measured device


characteristics for a
particular depletion mode
MOSFET

Fig. 7.20 The long RC time


constant for a capacitor
connected to the gate of a
MOSFET can be used to
make a two-state memory
element

by the RC time constant. With the extremely large resistance at the MOSFET gate,
Rgate , even a very small capacitance (a few pF or even less) can lead to an appreciable
time constant. In fact, the inherent capacitance of the gate connection may be large
enough to be useful without an additional capacitor. For the example of Fig. 7.20 the
state of the capacitor (charged or discharged) can be read at a later time by connecting
the drain to +V through a resistor. A low output value, near 0 V, indicates the capacitor
is charged, and a higher value, near +V, that the capacitor is discharged.
Memory that relies on the charged state of a capacitor will need to be refreshed
from time to time and is referred to as dynamic memory. The output is simply read
often enough, compared to the RC time constant, and the result is used to either
recharge or re-discharge the capacitor. The simplicity of the memory unit makes up
for the need to periodically refresh the memory when a large amount of memory is
being used.
152 7 FETs

7.8 Problems

1. In the circuit of Fig. 7.P1, the gate to source voltage is V gs = − 1.0 V. What is
the drain to source voltage, V ds ?
2. An n-channel JFET is characterized by the model parameters I dss = 5 mA and
V p = –3 V. Sketch, based on this information, the characteristic curves for V gs
= 0, –1 V, and –2 V.
3. The n-channel JFET shown in Fig. 7.P3 is characterized by model parameters
I dss = 5 mA and V p = –3 V. What are the gate to source voltage, V gs , the drain
to source voltage, V ds , and the drain current, I d for this circuit?
4. If the MOSFET characterized in Fig. 7.19 is used near an operating point of V ds
= 12 V and I d = 2 mA, what is the appropriate value for the transconductance,
gm ?

Fig. 7.P1 Problem 1

Fig. 7.P3 Problem 3

5. If the MOSFET characterized in Fig. 7.18 is used in the ohmic region, V ds <
0.5 V, and with V gs = 3.6 V, what is the effective drain to source resistance of
the device?
6. (a) Show that the JFET mathematical model can be written in the alternate form
( )2
Id = K Vgs − Vt .
7.8 Problems 153

(b) If I dss = 10 mA and V p = −3 V, what are the corresponding parameters, K


and V t ?
(c) Show, based√on this model and at an operating point with drain current I d ,
that gm = 2 K Id .

7. Though they might not be identified as such, introductory physics courses include
several results based on small changes. In hindsight, can you identify one (or
more) of these?
Chapter 8
Bipolar Junction Transistors

The bipolar junction transistor, or BJT, is a three-lead device made by stacking three
layers of doped semiconductor material with alternating p- and n-doping. Hence,
there are NPN or PNP type transistors. While the details of the electronic behavior
of these transistors differs from that seen for the FET, the general approach for
circuit analysis remains the same. Solutions are found for an operating point based on
device characteristics, that is, the “large signal behavior,” then, if necessary, linearized
models are used to look at small changes from that operating point. While the FET can
be considered a voltage-controlled device, the BJT is considered current-controlled.
For the BJT, the three leads are called the collector, base, and emitter. The basic
structure of an NPN transistor looks something like Fig. 8.1. While that figure gives
the appearance that the emitter and collector could be switched, in real transistors
the geometry is much less symmetrical and the performance with these two leads
switched is usually quite poor.
Schematically, BJT transistors are drawn as in Fig. 8.2, where the arrow is on the
emitter and points from “P” toward “N”.
A BJT in normal operation will have the base-emitter junction forward biased
and the collector–base junction reverse biased. To a good approximation, the basic
behavior of the BJT can be specified in many circumstances by the relationships
between I b , I c (or I e ), and V ce , without worrying about V be too much. The forward
biased base-emitter junction will result in a significant base current, I b , and a small
base to emitter voltage, V be .
Despite the fact that the collector–base junction will be reverse biased, the
collector current can be quite large due to the extra charge carriers that will be
present in the base region when the base-emitter junction is conducting. The motion
of these extra “injected” charge carriers is dominated by thermal diffusion. If the
base region is physically small enough, once those charges get into the base region,
there is a much better chance that they will diffuse to the collector rather than out
through the base. When there are no extra injected carriers the collector current will
be a reverse saturation current—essentially zero for most practical purposes.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 155
B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_8
156 8 Bipolar Junction Transistors

Fig. 8.1 The three-layer


structure of an NPN bipolar
junction transistor

Fig. 8.2 Schematic symbols


for NPN and PNP transistors

For the purposes of basic circuit analysis, however, only the electrical character-
istics need to be known, and it is less important why the characteristics are what
they are. Readers interested in learning more about semiconductor theory, and how
it applies to transistors, are referred to more advanced texts on the subject.
As was the case for the FET, the BJT has multiple voltages and multiple currents
that are relevant. As already mentioned, of typical concern will be the collector
current (similar to the drain current for the FET), the collector to emitter voltage
(similar to the drain to source voltage for the FET) and the current into the base. For
the FET, the gate (to source) voltage acted to control the drain current. For the BJT,
the base current acts to control the collector current.
The large signal characteristics for an NPN BJT can be measured using a circuit
such as shown in Fig. 8.3, where V be , I c , and I b are measured, for various values of
V ce , resulting in a family of curves. The measured results for a particular transistor
are shown in Fig. 8.4. The curves on the left (8.4a) are the characteristic curves
analogous to those for the FET. When V ce is larger than about 1 V, the transistor is in
the active region. When V ce is smaller than about 1 V, the transistor is in the saturation
region. As was the case for the JFET, the active region is used for amplifiers. In the
saturation region, both junctions may be forward biased causing the transistor to
behave like a wire. The saturation region is used for on/off applications including
basic switching and some digital electronics.
The measured curves on the right (8.4b) show that in the active region V be does
indeed appear to be a “turn-on voltage of about 0.5 V” (in this case, close to 0.7 V) over
a wide range of conditions. While these curves are for a particular NPN transistor,
similar behavior can be expected for other NPN transistors. For a PNP BJT, add a
minus sign to all voltages and currents in these curves.
8.1 BJT D.C. Model 157

Fig. 8.3 Basic circuit used to define transistor characteristics. Shown for NPN. For PNP, all currents
and voltage sources are reversed

Fig. 8.4 Measured transistor characteristics for a small-signal NPN transistor

8.1 BJT D.C. Model

The “transfer characteristic” refers to I c as a function of I b , for different values of


V ce . That is, for selected fixed values of V ce the data is read off the previous graph
and those points are connected with a smooth curve. Figure 8.5a shows the transfer
characteristics corresponding to the BJT characteristics shown in Fig. 8.4. Notice
that to a reasonable approximation, and over a wide range of values for V ce (but
always in the active region) the transfer characteristic is described approximately by
a straight line and it is almost the same line over a wide range of values for V ce .
Hence, as a model, it is possible to write I c = βI b for some constant β. Such a line
is shown in Fig. 8.5b for this transistor, with β ≈ 250. This value for β would be
referred to as the “DC current gain.”
Since β is the ratio of two currents, it does not need any units. Typical values for
β are 30–300. Hence, approximations that result when β >> 1 can be used routinely.
Note also that β can vary by as much as a factor of 2 from one transistor to another
of the same type, so it is rare that highly accurate computations will be necessary.
Measurements of V be for this same transistor as a function of base current, I b ,
for different values of V ce , are shown in Fig. 8.4b, and are also simple to model.
158 8 Bipolar Junction Transistors

Fig. 8.5 Collector current as a function of base current for a measured values for several different
collector-emitter voltages and b for a simple model that approximates the measured values

Fig. 8.6 A simple model


that can be used to
approximate the data of
Fig. 8.4b

For the active region, there is very little dependence on V ce . Hence, the extremely
simple model of Fig. 8.6 might be appropriate—that is a constant voltage (~0.7 V
in this case) such as from a battery. The data shows a slight slope upward for higher
currents. If that trend is important, then the model can also be given a slight upward
tilt (i.e., a small series resistance).
Combining these ideas, a simple model for a BJT made from linear components,
valid in the active region, would look like those of Fig. 8.7, where r b is a small-valued
resistor and r c a large-valued resistor. The battery provides the few tenths of a volt
appropriate for V be , but of course it is not a real battery and cannot supply power.
Here the battery is shown as 0.5 V, a convenient value to remember and use. The
dependent current source supplies a collector current proportional to the base current.
As is the case for the battery, it is not a real current source and cannot supply power
on its own.
8.1 BJT D.C. Model 159

Fig. 8.7 Simple, linearized models for BJTs when used in the active region

Fig. 8.8 Extremely simple


models for BJTs. These
models can be used if the
resistors and sources shown
in Fig. 8.7 are negligibly
small compared to other
components in the circuit

The small-valued resistor, r b , is often optional, and provides a small upward slope
on the line as seen in the measured values of V be versus I b . The larger resistor, r c ,
plays the same role as did r d for the JFET, and allows for the fact that the characteristic
curves are not ideally flat. In many applications it can be omitted. The convention
being followed here is that lower case r’s are used for resistors within the model,
reserving upper case R’s for real resistors.
In many cases, a circuit might be understood with an even simpler model for the
transistor. The battery voltage is taken to be 0 V, r b = 0 (a wire), and r c → ∞ (an
open circuit) resulting in simplest BJT models shown in Fig. 8.8.
A first look at some of the following transistor problems might lead one to think
that neglecting components is routine and is done in a somewhat cavalier manner—
that is, using a technique sometimes referred to as hand waving. Whether or not any
components of any model can be neglected depends, of course, on the values of other
components in the circuit. Hence the very simple models should always be used with
caution. The examples below will illustrate how such simplifications can be justified
based on the relative size of component values. Whether or not such simplifications
are valid must be considered on a case-by-case basis and should always be justified,
or at least, justifiable.
Note that when the simple models work, the FET and BJT can be described
similarly except that the FET is voltage controlled (by V gs , with negligible gate
current) and the BJT is current controlled (by I b , with small/negligible base-emitter
voltage). In both cases an “output current” (I d for the FET, I c for the BJT) is what is
being controlled.
The appropriate model parameters may depend on how the transistor is being
used. The models above are motivated by the large signal behavior. For the BJT
those models can be used to find the operating point—that is the results of the large
160 8 Bipolar Junction Transistors

signal analysis in the active region. Of course, the graphical method, as was used for
the FET, can be used for BJTs as well.

8.2 BJT A.C. Model

When used for an amplifier, the same basic BJT model can be used, though with
parameters appropriate for small changes in the voltages and/or currents. To consider
small changes from an operating point, proceed as was done for the JFET. The result
is that the BJT transistor model will look the same as above (Figs. 8.7 and 8.8), but
the values of the parameters may need some adjustment.
When considering small changes from an operating point, the slope of the transfer
characteristic near the operating point is considered. That is,
| |
d Ic || /\Ic ||
β= ≈ , (8.1)
d Ib |Vce constant /\Ib |Vce constant

which can depend on the operating conditions. If the transfer characteristic is a


straight line, this value is the same as the DC current gain. It is often the case that
the DC and small signal current gains can be taken to be the same—they are close
enough in value—and at the risk of some confusion, the same symbol, β, is used
for both. In some cases, the change in the slope of the transfer characteristic may
be important. For the transfer characteristic shown above near I b = 100 µA, β (for
small changes) will be somewhat larger for V ce = 15 V than for V ce = 3 V. More
complicated models for these transistors refer to “h-parameters” (hybrid parameters),
of which there are many, and the h-parameter corresponding to β is “hfe .”
For an idealized BJT, the characteristic curves in the active region would be flat.
In cases where the slope is important, a parameter can be used to describe the slope
of those lines,
| |
1 d Ic || /\Ic ||
= h oe = ≈ , (8.2)
rc d Vce | Ib constant /\Vce | Ib constant

evaluated at a point on the curves near where the transistor is being used. Typically,
r c ≈ 1 k to 10 k. Aside from some name changes, this is identical to what was done
to find r d for the JFET model.
Another parameter that is sometimes important, and which may not be evident in
the characteristic curves, relates the voltage across the base to emitter junction to the
resulting current into that junction, I b . That junction looks somewhat like a diode,
so an appropriate diode model can be used. It is convenient for simple calculations
to use a piece-wise linear model such as

Vbe = Vbe0 + Ib rb , (8.3)


8.2 BJT A.C. Model 161

where V be0 is the “turn on voltage” for the junction, which is about 0.5 V, and
|
d Vbe ||
rb = h i = , (8.4)
d Ib |Vce
e
constant

evaluated near the operating point. Typically, r b ≈ 10 to 1000 y, and the value
depends strongly on the operating point. The constant DC value, V be0 , will not be
important for AC signals.

8.2.1 BJT Large Signal Example—Graphical Solutions

For the following example, the transistor has the characteristics shown in Fig. 8.9a.
The open circles indicate data points that are read off of the characteristic for V ce =
7 V to create the transfer characteristic, Fig. 8.9b. That value of V ce was chosen only
because it is near the middle of the active region shown. For reference, several other
such curves for different choices of V ce are shown with the dashed lines. The dotted
line is discussed below.
Now consider such a transistor in the circuit of Fig. 8.10. What is the collector
current I c , and the collector to emitter voltage, V ce ? Remember that all grounds are
connected together and are considered to be the zero-volt reference.
Applying Kirchhoff’s voltage law from the 12 V supply, through the transistor
and out the emitter generates the so-called “load line” for this configuration.

12V − 200Ic − Vce − 10(Ic + Ib ) = 0. (8.5)

Fig. 8.9 The transistor characteristics shown in a are used to create the transfer characteristics in
b as part of a graphical solution to a BJT circuit
162 8 Bipolar Junction Transistors

Fig. 8.10 An example BJT


circuit to be solved

Similarly, going through the 1 V supply and the out the emitter, generates the
so-called “bias line,”

1V − 1kIb − Vbe − 0.01k(Ic + Ib ) = 0. (8.6)

There are too many unknowns to be able to solve these equations by themselves.
The transistor properties need to be included in some way, at least approximately.
As a very simple approximation, as mentioned above, take V be ≈ 0.5 V to get

0.5V − 1.01k Ib = 0.01k Ic , or Ic = 50mA − 101Ib . (8.7)

This is a straight-line relationship between I b and I c and is plotted on the transfer


characteristic (see the dotted line in Fig. 8.9b). While a value for V ce is still unknown
at this point, it is clear from the graphical solution that, at least if 3 V < V ce < 9 V,
the current I c must be very close to 34 mA and I b will be close to 0.16 mA. Putting
these values back into the first equation gives

Vce = 12V − 0.2k · 34 mA − 0.01 k · 34.16 mA ≈ 12 V − 0.21 k · 34 mA = 4.9 V.


(8.8)

This value of V ce is indeed within the active range and this solution should be
reasonable.
The two important properties of the BJT that made this solution possible are that
within the active region V be is almost constant and that the transfer characteristic
does not depend on V ce very much.

8.2.2 Single Supply Operation

In the example above, two sources of power are shown—one a 12 V supply and the
other a 1 V supply. In practice this is inconvenient and rarely necessary. Instead, the
8.2 BJT A.C. Model 163

Fig. 8.11 Thevenin’s


theorem is used to create an
equivalent circuit to that
shown in Fig. 8.10 which
requires only one voltage
source

single larger supply is used, along with Thevenin’s theorem, to create an equivalent
circuit.
Consider the circuit shown in Fig. 8.11. The two “biasing resistors,” R1 and R2 ,
are to be chosen so that the circuit is equivalent to that of Fig. 8.10. A Thevenin
equivalent (refer to Chap. 2) is created for the biasing circuit, as shown on the right,
where
R2
Vb = 12V
R1 + R2 (8.9)
Rb = R1 ||R2 .

Solving to match the values in the example above, that is V b = 1 V and Rb =


1 k, yields R1 = 12 k, R2 = 1.1 k. When solving single-supply circuits, a first step
is often the reverse process—to create the Thevenin equivalent two-supply circuit
from a pair of biasing resistors.

8.2.3 Solutions from Parameters

It is often the case that full transistor characteristics are not available and solutions
must be obtained using limited parameters that have been supplied. For the transistor
in the example above, the transfer characteristic shows that β ≈ 200. Consider that
same transistor in the circuit in Fig. 8.12, assuming the value of β is all that is known
about the transistor.
As a first step in this analysis, replace the biasing circuit (the resistors on the left
plus the power supply) with its Thevenin equivalent (Fig. 8.13). Now use the transistor
model, which assumes the active region, to get the linear circuit of Fig. 8.14. Using
Kirchhoff’s voltage law around the left and right loops,

1.1V − Ib (0.91k + rb ) − 0.5V − (Ic + Ib )(0.1k) = 0


(8.10)
12V − Ix (0.5k) − Vce − (Ic + Ib )(0.1k) = 0.
164 8 Bipolar Junction Transistors

Fig. 8.12 A second example


BJT circuit to be solved

Fig. 8.13 The first step to


solve the circuit of Fig. 8.12
is to use Thevenin’s theorem
for the biasing resistors

Fig. 8.14 The second step is


to replace the BJT with the
linearized model

Here V ce equals the voltage across the current source. Since r c will be significantly
larger than 500 y, r c will have a much smaller current through it and can be neglected
in comparison (see Problem 6 of Chap. 1). With that approximation, I c = βI b . The
equations become

(1.1 − 0.5)V − Ib (0.91k + rb + (1 + β)(0.1k)) = 0


(8.11)
12V − Vce − Ib (β(0.5k) + (1 + β)(0.1k)) = 0.

In the first equation, with β = 200, the 100 y resistor acts like it is close to 20 k,
and since r b << 10 k, r b can be neglected in comparison. Then the first equation
can be solved for I b (and hence also I c = βI b ) and then that result is put into the
8.3 BJT Amplifiers 165

Table 8.1 The β dependence


β I b (µA) I c (mA) V ce (V)
for the operating point of the
example calculation 50 100 5.0 9.0
100 54 5.5 8.8
200 29 5.8 8.5

second equation to find V ce . For this example, I b = 29 µA, I c = 5.8 mA, and V ce =
8.5 V. Since V ce > 1 V, this is in the active region, as was assumed above, and so the
solution should be reasonable.
Using this example, it is possible to estimate some of the effects of the inevitable
variability from transistor to transistor. Solving the equations above with β = 50,
100, and 200 yields the results shown in Table 8.1. Notice that I c and V ce do not
change very much. If the equations are solved leaving β as a variable, then

β(0.6 V)
Ic = , (8.12)
(1 + β)(100 y) + 910 y

and it can be seen that I c will be roughly independent of β provided β >> 1, which
is typical, and also if β(100 y) >> 910 y, which is certainly valid for this example.
One good reason to include the ~100 y emitter resistor in such a circuit is to mini-
mize some of the effects of the significant variations in β that occur from transistor
to transistor. As will be seen later, the emitter resistor also can have a significant
impact on amplifier gain.

8.3 BJT Amplifiers

There are three basic configurations for single-transistor BJT amplifiers, shown for
NPN transistors in Fig. 8.15. The common base configuration is usually drawn so
that the input is on the opposite side of the drawing from the output. Here it is drawn
to facilitate comparison between the three amplifiers. Note that the “large signal
analysis” or “DC analysis” to find the operating point is identical for all of these.
That large signal analysis was shown above.
Consider the Common Emitter (CE) amplifier in some more detail. The operating
point is found as before and it is assumed that this has been done and the transistor is
indeed in the active region. The next step is to look at the AC small signal analysis.
The overall procedure here is identical to what was done for the FET transistor
amplifiers; however, some important details have changed.
To keep matters simple, consider time dependent input signals, and hence also
output signals, that have a frequency large enough so that the impedance of the
input and output capacitors is negligible (or alternatively, that “the capacitors are
large enough”). Such an assumption is certainly not necessary and is not always
166 8 Bipolar Junction Transistors

Fig. 8.15 The three single-transistor amplifier configurations used with BJTs. Shown for NPN
transistors

appropriate. As was the case for the FET amplifiers, it is done here to keep the
analysis simple. With that assumption, those capacitors can be replaced with wires
for this part of the analysis. In the case of the CE amplifier, the emitter capacitor
may or may not be present and so the more general situation is treated here—the
combined emitter resistor and capacitor are treated as an impedance Z e .
Now the transistor model is used with all DC voltage sources “turned to 0 V” (i.e.,
replaced with wires to ground). Putting all the 0 V connections (the grounds) at the
bottom of the figure, the common emitter amplifier now looks, for the AC signals,
as shown in Fig. 8.16. Remember that the voltages and currents here refer only to
changes from the operating point and when one is considering changes, any constant
voltage is the same as zero—after all, there is zero change for a constant voltage.
Kirchhoff’s law from the input and down through the emitter gives

Vin − Ib rb − (Ib + Ic )Z e = 0. (8.13)

Kirchhoff’s law around the right loop gives

−Ic Rc − (Ic − β Ib )rc − (Ib + Ic )Z e = 0, (8.14)

which can be solved to get

Fig. 8.16 The BJT model is


used to determine the voltage
gain for small AC signals.
Note that it is assumed the
transistor is in the active
region. Note also that all of
the DC voltages are 0 VAC
8.3 BJT Amplifiers 167

βrc − Z e βrc
Ic = Ib ≈ Ib ≈ β Ib , (8.15)
r c + Rc + Z e r c + Rc + Z e

where the first approximation will be very good in virtually all circumstances and
the last approximation is valid if r c is large enough compared to the collector and
emitter resistors, which is often, though not always, true. Substituting back into the
first equation, and noting that V out = −I c Rc , yields

β Rc r c β Rc
Vout = −Ic Rc = −Vin ≈ −Vin ,
rc (rb + (1 + β)Z e ) + Rc (rb + Z e ) rb + (1 + β)Z e
(8.16)

where the approximation is certainly valid if r c is large compared to Rc . Consider


also the case where there is no emitter capacitor, and βRe >> r b . Then V out /V in ≈
−Rc /Re . On the other hand, if an emitter capacitor with a large valued capacitor is
included, so the total emitter impedance is very small at the frequencies of interest,
then β|Z e | << r b , so V out /V in ≈ −βRc /r b . The advantage of the former (no emitter
capacitor) is that the voltage gain is independent of β, and hence is much more under
the designer’s control. The advantage of the latter is that the gain is much larger. Do
recall however, that the emitter resistor, which is part of Z e , is very important for
the DC analysis, the “biasing,” and so that resistor cannot be simply omitted to get
a large voltage gain—hence the use of a bypass capacitor to short the AC signals to
ground without affecting the DC behavior.
Also of interest for an amplifier is knowledge of its input and output impedances.
That is, what does the amplifier look like to the other circuits that are connected to
it? For the amplifier above, if |Z e | is small enough so that it can be replaced by a
wire, the input impedance is obviously (R1 ||R2 )||r b which will be relatively small
since r b is generally small. On the other hand, if there is no emitter capacitor, the
calculation is somewhat more complicated due to the fact that the current source is
now a dependent current source. The superposition principle, where each source is
considered independently, does not apply for dependent sources.
To do this calculation, use the fact that the input impedance, Z i , is given by Z i =
V in /I in and
( )
Vin 1 r c + Rc + Z e
Iin = + Ib = Vin + ,
R1 ||R2 R1 ||R2 rc (rb + (1 + β)Z e ) + rb (Rc + Z e )
(8.17)

to get the very complicated and non-intuitive result,


( )−1
1 r c + Rc + Z e
Zi = + , (8.18)
R1 ||R2 rc (rb + (1 + β)Z e ) + rb (Rc + Z e )

and so if |Z e | is very small, the expected result is obtained. If that is not the case, the
situation is more complicated. Looking at the case where both r c and β are large,
168 8 Bipolar Junction Transistors

then

Z i = (R1 ||R2 )||(rb + (1 + β)Z e ). (8.19)

In the case where there is no emitter capacitor it is often true that (1 + β)Re >>
r b , due to the large values of β, so that r b can be neglected.
The output impedance, Z o , is the Thevenin equivalent impedance at the output of
the amplifier. If the dependent current source depends only on the input voltage, but
not the output voltage, it is possible to use the superposition principle and “look back
into the circuit with the sources set to 0” to find the Thevenin equivalent resistance.
For the common emitter amplifier with a large emitter capacitor, so |Z e | is small, the
latter method works and the output impedance is then Rc ||r c ≈ Rc if r c is large. If
|Z e | is not negligibly small, however, more care is necessary. In that case, the current
source depends on the voltage at the output even if the input voltage is “turned off.”
Recall that the Thevenin impedance is the ratio of the open circuit voltage and the
short circuit current. The open circuit output voltage, V out , was computed above (see
8.16). The AC model with the output connected to a short circuit looks like Fig. 8.16
but with Rc = 0, and so the short circuit output current, I os , is I c , as computed above,
but with Rc = 0. Then

Vout Rc rc (rb + (1 + β)Z e )


Zo = = , (8.20)
Ios rc (rb + (1 + β)Z e ) + Rc (rb + Z e )

and it can be seen that if r c and/or β is large enough so that the second term in the
denominator is negligible, then Z o ≈ Rc . If the output capacitor, C o , is not negligible,
its impedance will need to be added to the above result.
Now for some more concrete examples.

8.3.1 Example—Common Emitter Amplifier

Estimate the AC voltage gain for the circuit in Fig. 8.17 that uses an NPN transistor
with β = 36.
First, assume the transistor is operating in the active region—that assumption will
be checked along the way.
Now do the DC analysis with the input signal off. At DC all the capacitors look
like open circuits and can be removed from the analysis. Simplify the biasing on
the left using the Thevenin equivalent as shown in Fig. 8.18, and put this into the
transistor model to get the linear circuit in Fig. 8.19. Now one can hope to neglect
r c (remove it) since typically r c > 1 k and the 100 and 50 y resistors are both much
smaller than 1 ky.
Using Kirchhoff’s voltage law,
8.3 BJT Amplifiers 169

Fig. 8.17 A common


emitter amplifier to be
analyzed as an example

Fig. 8.18 The first step to


analyze the amplifier of
Fig. 8.17 is to use Thevenin’s
theorem to simplify the
biasing circuitry

Fig. 8.19 The second step


to analyze the amplifier of
Fig. 8.17 is to assume the
active region and put in the
(DC) transistor model. Here
the capacitors are treated as
open circuits

4V − (1.2k + rb )Ib − 0.5V − (1 + β)Ib · 50y = 0, (8.21)

so with β = 36 and assuming r b is small enough to be negligible (here that means


r b << 1.2 k),

Ib = 3.5V/3k = 1.17mA, (8.22)


170 8 Bipolar Junction Transistors

Fig. 8.20 The last step to analyze the circuit for small AC signals. Here the capacitors are approx-
imated as short circuits. If the frequency of the signal is not high enough, then the impedance of the
capacitors must be included

so I c = 42 mA giving V ce = 10 V − (42 mA)(100 y) + (43 mA)(50 y) = 3.6 V.


That is in the active region (i.e., V ce > 1 V), validating the initial assumption.
Now look at the AC (i.e., time-dependent) model. For simplicity, assume the
capacitors (with unknown values) are large enough so that they can be treated as
short circuits. Then, realizing that the DC voltages are “0 V AC” the equivalent
circuit for the time-dependent signals is as shown in Fig. 8.20. Then I b = V i /r b and
neglecting r c (it is large compared to 100 y on one side and shorted to ground at the
emitter), I c = 36I b and V out = 100I c . This gives

Vout / Vin = (36/rb ) · 100y, (8.23)

and to get a value, r b needs to be known. Note that if r b had been neglected here, the
result would have been unusable—in this case r b is not “small enough compared to
other resistors” since there are no others between the input and ground.
Using the exponential model for the base to emitter p–n junction, the resistance
is given (at room temperature) roughly by1

rb = 0.026V/Ib = 0.026V/1.17mA = 22y. (8.24)

where I b is the value at the operating point (from the DC analysis). Hence a voltage
gain of about 160 should be expected.

8.3.2 Example—Common Collector Amplifier

Estimate the AC voltage gain for the circuit in Fig. 8.21 that uses an NPN transistor
with β = 36.

1 The exponential model for diodes is in Chap. 6. The resistance for small changes is found from
the derivative, dV d /dI d = 1/(dI d /dV d ). Such an estimate is accurate only to factors of order unity.
Many manufacturers supply graphs from which one can find r b = hie for specific transistors under
a variety of operating conditions.
8.3 BJT Amplifiers 171

Fig. 8.21 A common


collector amplifier analyzed
as an example. The biasing
for this example is the same
as for the example circuit of
Fig. 8.17

This is the same as Example 1, except using the common collector configura-
tion. Hence, the large-signal DC analysis is identical to Example 1 and need not be
repeated.
Again, for simplicity assume the capacitors act as shorts for the time depen-
dent signals. At lower frequencies, the impedance of the capacitors will need to be
considered.
Putting in the AC model for the BJT and again noting that the constant 10 V
source is 0 VAC, the AC model for the circuit is as shown in Fig. 8.22, and so

Vin − Ib rb − (Ib + Ic ) · 50 = 0
β Vi (8.25)
Ic ≈ β Ib = ,
rb + (1 + β) · 50

and if (1 + β) 50 >> r b , which is certainly the case here, then

Vout = (Ic + Ib ) · 50 = Vin . (8.26)

The voltage “gain” is very close to 1. That is, the output voltage “follows” the
input voltage.

Fig. 8.22 The transistor


model is placed into the
circuit of Fig. 8.21 to find the
AC small signal gain
172 8 Bipolar Junction Transistors

Fig. 8.23 The PNP common emitter amplifier at a is derived from the NPN amplifier by replacing
the transistor and reversing the power supply. The circuit is more commonly shown with the higher
voltage at the top as in b, for a negative supply, or c, for a positive supply

This amplifier has a low output impedance (here, 50 y) whereas the input
impedance is about β times larger. Hence there is a large current gain, and hence a
large power gain, even though there is little change in the voltage.
In summary, the same basic procedure is used for all transistor amplifiers. First,
the large signal analysis with no input is used to find the “operating point.” Then
parameters for a suitable model are found to describe small changes about that
operating point. The model is kept simple and linear. The model is placed in the
circuit to look at the changes that occur (e.g., for an AC signal). Along the way,
various approximations might be possible to simplify the analysis. At each step, of
course, those approximations need to be justified.
The examples above use NPN transistors. If a PNP transistor is used, the voltages
and currents will be reversed. An example is shown for the common emitter amplifier
in Fig. 8.23a. It is common practice to place larger voltages toward the top of the page,
so the common emitter amplifier with a PNP transistor will appear as in Fig. 8.23b
if a negative power supply is used and as in Fig. 8.23c if a positive supply is used.
Remember that only the difference in voltage matters.

8.3.3 Using the Saturation Region

Transistors are also used as on/off devices, both for switching and for digital logic.
When a BJT is used in this way, the transistor model does not apply. However, that
model can be used to estimate the results if the model is used carefully.
Example 1 Consider the circuit of Fig. 8.24 where V in is either 0 V or 5 V (relative
to ground). Assume that for this transistor, β ≈ 250. Going around the input loop
Kirchhoff’s voltage law yields.
8.3 BJT Amplifiers 173

Fig. 8.24 A very simple


single transistor circuit for
on/off applications

Vin − 10kIb − 0.5V = 0. (8.27)

so I b = 0 or 0.45 mA depending on V in .
When I b = 0, then I c = 0 and V out = 5 V − 1 k I c = 5 V. In this state the transistor
is “off” and the output sees only the 5 V power source through the 1 k resistor.
When I b = 0.45 mA, the model is used to compute I c = 250 and I b = 112.5 mA.
Then V out = 5 V − 1 k I c = − 107.5 V. Since the transistor cannot supply power, it
merely reroutes power, it must be the case that 0 V < V out < 5 V and so this solution
is clearly invalid. The output will, in some sense, get as close to this solution as it
can while still being within the allowed range. Hence the solution is very close to
0 V.
What happens can be seen graphically by plotting the load line (V ce = 5 V −
1 k I c ) on the characteristic curves (Fig. 8.25). The solution based on the model is
negative, but the model does not apply on the negative side of the V ce axis. The real
transistor characteristic gives a solution in the saturation region with a very small
voltage for V ce . A small voltage with a large current is similar to the behavior of a
wire. In that state the transistor is said to be “on”.
This particular circuit takes an input voltage of 0 V or 5 V and produces a corre-
sponding 5 V or 0 V as the output. In a sense, the circuit provides “the opposite of
the input.”

Fig. 8.25 A graphical


solution for the circuit of
Fig. 8.24
174 8 Bipolar Junction Transistors

Fig. 8.26 An example of a


transistor being used as an
on/off switch for a light bulb

Example 2 The saturation region is also used to switch higher power devices using
a low power control signal. As an example, consider a 12 V, 55 W light bulb that is
to be switched using an NPN transistor and a 5 V signal (Fig. 8.26).
If the bulb is specified as a “12 V” bulb and a 12 V power supply is used, that
means to achieve full brightness, virtually all of the 12 V available must be across
the bulb, and not the transistor. Hence, to turn on the bulb (to full brightness), V ce
must be close to 0 V. If it is a 55 W bulb, then the current can be computed:

P = V I → I = P/ V = 55W/12V = 4.6A. (8.28)

A transistor must be chosen which can handle that much current (or more). Such
information is available on the manufacturer’s transistor data sheet. Also on that data
sheet will be the expected range for the DC current gain, β. To be safe, the minimum
expected value should be used for calculation. For use as an example, assume that,
for the transistor chosen, that minimum is β = 50.
Then to turn on the light, you need I b > 4.6A/50 = 92 mA. If the signal, V b =
5 V, then solving for R,

5V − Ib R − 0.5V = 0, (8.29)

so R < 49 y.
The strategy here was to use β, a parameter for the active region, to estimate the
values necessary to end up in the saturation region.

8.4 Problems

1. For the transistor characteristics shown in Fig. 8.4a, estimate the AC current gain
and the value of the model resistor r c near an operating point with V ce = 7 V and
I c = 15 mA.
8.4 Problems 175

2. Derive an expression for the voltage gain of the common base configuration
using the transistor model and appropriate approximations. Be sure to justify
any approximations. Assume the biasing has already been done and that the
transistor is in the active region.
3. The circuit in Fig. 8.P3 can provide a current through a load, RL , which is almost
constant as the applied voltage, +V, changes, provided that voltage is within a
certain range. If the transistor has β = 150 and the desired applied voltage range
is 5 V < +V < 15 V, what current, I, is expected and what is the maximum value
that can be used for RL ?

Fig. 8.P3 Problem 3

4. Electric Eye: For the circuit in Fig. 8.P4, the NPN transistor is specified as having
hfe = β = 75. The resistor RL is a model for a 12 V, 1 W light bulb (i.e., the
light bulb is designed to be used with 12 V across it—significantly more and it
will burn out, significantly less and it will not be at full brightness. Significant
here is approx. ±10%). The variable resistor, Rs , is a photoresistor2 where the
resistance when exposed to a light beam is 100 y and when the beam is blocked,
it is 100 ky. Estimate an appropriate value, or range of values, for Rx so that the
light bulb is (nearly) fully on when the light beam is blocked, and is (essentially)
off when the beam is unblocked.
5. Solve for all the currents and for V ce for the PNP transistor circuit shown in
Fig. 8.P5 assuming the transistor has β = 150.

2 A photoresistor is a bit of semiconducting material that can be exposed to light. If the photons of
light have an energy larger than the semiconductor band gap, electrons and holes can be created
by the light. As the concentration of these charge carriers increases, the resistance of the device
decreases.
176 8 Bipolar Junction Transistors

Fig. 8.P4 Problem 4

Fig. 8.P5 Problem 5


Chapter 9
More on Amplifiers

The previous chapters showed, in some detail, how to make amplifiers using a single
transistor. Further detailed analysis for any particular amplifier is left to those who
find it necessary. Here some additional ideas applicable to amplifiers are presented
more generally. To start, yet another general theorem for linear circuits is considered.
After that, some useful configurations that rely on multiple transistors are presented,
as well as some more general concepts that apply to any linear amplifier.

9.1 Miller’s Theorem

While the inner workings of an amplifier may involve non-linear devices, the ampli-
fiers shown can be treated as a device that is linear—at least as long as the input and
output signals do not become too large or go out of range. The basic device rule for
a linear amplifier is that there is an output voltage (or current) that is proportional to
the input voltage (or current).
Miller’s theorem is useful for linear amplifiers, in particular for understanding the
origin of the upper frequency limits of many amplifiers. Miller’s theorem can also
be used for some circuit analysis when feedback is employed. Feedback will play a
very important role for the op-amp circuits discussed in the next chapter.
To start, consider a box (containing electronics) that has three connections where
one of them is the zero-volt reference (i.e., ground). The other two are labeled “1” and
“2.” The current out of these connections is I 1 and I 2 respectively, and the voltages
measured relative to ground are V 1 and V 2 . Miller’s theorem applies if the box has
the property that V 2 = kV 1 , where k is some constant. Now consider such a box with
an impedance Z connected between 1 and 2 as shown in Fig. 9.1a.
It is a simple matter to compute the relationships between I 1 and V 1 , and I 2 and
V 2,

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 177
B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_9
178 9 More on Amplifiers

Fig. 9.1 Miller’s theorem shows that the circuitry inside the box will be unchanged if the circuit
in (a) is converted to the circuit in (b)

V1 − V2 V1 (1 − k) V1
I1 = = =
Z Z (Z /(1 − k))
V2 − V1 V2 (1 − 1/k) V2
I2 = = = , (9.1)
Z Z (k Z /(k − 1))

and, so far as the electronics within the box is concerned, the original configuration
is equivalent to attaching an impedance Z 1 from connection 1 to ground and an
impedance Z 2 from connection 2 to ground, provided the values

Z kZ
Z1 = ; Z2 = , (9.2)
1−k k−1

as shown in Fig. 9.1b.


To see how this applies to understanding the frequency dependence of a transistor
amplifier, consider the “hybrid pi” BJT transistor model shown in Fig. 9.2a and a
similar model for a JFET in Fig. 9.2b. These models are a bit more sophisticated
than what was used previously and include some of the capacitance associated with
the p-n junctions that had been previously ignored.
Figure 9.2 includes typical ballpark values for a low-power transistors. Capaci-
tance values for higher power transistors can be much larger. If the BJT transistor

Fig. 9.2 A more detailed model for a a BJT and b an FET transistor that includes some of the
internal capacitance
9.2 Two-Transistor Configurations 179

Fig. 9.3 Miller’s theorem


applied to the BJT in Fig. 9.2
when the transistor is used in
a common emitter amplifier

shown is used in a common emitter amplifier, which has a large voltage gain, the
small collector to base capacitance of roughly 3 pF is directly connected between the
input (the base) and the output (the collector). If the voltage gain of this amplifier is
−A, Miller’s theorem shows this is equivalent to connecting a capacitance from the
input to ground with a value (1 + A) larger, illustrated in Fig. 9.3. With the common
emitter configuration, it is easy to achieve values such as A ≈ 100. Hence there may
will be a significantly larger effective capacitance connected between the input and
ground. That capacitance will short out higher frequency input signals before they
can be amplified.
The larger the amplifier gain, the larger is this effective capacitance between the
input and ground, and hence the lower the maximum frequency that will be amplified.
Similar behavior occurs in all linear amplifiers. A result is that the product of the
gain and the bandwidth (in this case the bandwidth is the upper frequency) is roughly
constant. If you try to vary the bandwidth or the gain, the other changes as well to
keep the product constant. In fact, the frequency response of many amplifiers will
be specified using the “gain-bandwidth product” and/or the “unity gain bandwidth”
for this reason. For example, if an amplifier circuit has a unity gain bandwidth of
100 kHz and it is used with a gain of 4, the bandwidth would be expected to be about
25 kHz.

9.2 Two-Transistor Configurations

There are several interesting configurations involving two (or more) transistors that
can accomplish a task that a single transistor cannot. Several of these are illustrated
in this section.

9.2.1 The Cascode Configuration

To make a (transistor) amplifier that works over a wider range of frequencies, the
capacitance of the device and the results of Miller’s theorem (above) must be taken
into account. One way to improve the high frequency response is to use a pair of
transistors configured as a cascode pair. For BJT transistors, this is essentially a
common emitter amplifier directly connected to a common base amplifier, as shown
180 9 More on Amplifiers

Fig. 9.4 The cascode


configuration uses two
transistors in place of one to
reduce the effects shown in
Fig. 9.3

in Fig. 9.4a. For the JFET, it would be a common source connected to a common
gate amplifier, as shown in Fig. 9.4b.
The common base and common gate amplifiers have a low input impedance and
hence tend to short the voltage gain (A) from the first transistor. Hence, the first
transistor has a small voltage gain, and thus a small Miller effect with a large current
gain. The capacitance between the output of the second transistor and the input is
very small and hence the Miller effect is greatly reduced.

9.2.2 The Darlington Pair

One way to increase the current gain beyond that of a single BJT is to create a
“Darlington pair,” as shown in Fig. 9.5. Ignoring 1 in comparison to β 1 , and β 2 (e.g.,
(1 + β i ) ≈ β i ), the effective current gain is now β 1 × β 2 which can be very large
indeed. Most often the reason to do this is not to get large currents at the output, but
to have small currents at the input—that is, a large input impedance.
Of course, the use of β to describe the transistor(s) assumes they are in their active
region. Maintaining that condition is much more difficult with the Darlington pair
and virtually impossible if one extends this idea to three or more transistors.
The Darlington pair works because the output of the first transistor, a current, is
an appropriate input for the second transistor. With FETs, which use a voltage to
control a current, such a pairing is problematic.

Fig. 9.5 A Darlington pair


of transistors can be used to
get a very large current gain
9.2 Two-Transistor Configurations 181

Fig. 9.6 The


complementary, or
“push–pull,” amplifier uses
one transistor for a positive
input, and a second for a
negative input

9.2.3 Complementary Symmetry Amplifier (“Push–Pull”)

One problem with the single transistor linear amplifiers is that they use a lot of power
even when there is no signal. Another problem is that they do not work well for
constant (DC) signals. Both of these effects are a result of the need for biasing—the
need to stay in the active region away from the origin of the device characteristics.
One way around this is a two-transistor amplifier known as the complementary
symmetry amplifier or “push–pull” amplifier. An example of such an amplifier made
using BJT transistors is shown in Fig. 9.6. Similar circuits can be produced using
FETs—in particular MOSFETs.
The biasing resistors are such that when there is no signal, both transistors are
just barely “off.” If a positive signal is present, the upper transistor conducts and
acts as an amplifier. The lower transistor is still off and does not do anything. On the
other hand, if a negative signal is present, the situation is reversed. Note that there
is a problem with this amplifier if the load is disconnected. During operation only
one transistor is on and is trying to send current to the load, however, if the load is
absent, the only place for the current to go is through the other transistor, which is
off. If no precautions are taken, this can, in some circumstances, cause the ratings of
one or the other transistor to be exceeded and possibly in the destruction of one or
the other transistor.

9.2.4 Differential Amplifier

The differential amplifier (or difference amplifier) has two inputs and the output is
proportional to the difference between the two inputs. For schematics, it is common
182 9 More on Amplifiers

Fig. 9.7 A general


schematic for a differential,
or difference, amplifier

Fig. 9.8 The two-transistor


configurations for a
difference amplifier based on
a BJTs and b JFETs

to use a triangle to indicate an amplifier if the internal contents are unknown or not
of immediate importance. Using that symbol, the differential amplifier is as shown
in Fig. 9.7.
The basic two-transistor differential amplifier configuration for BJTs and for FETs
is shown in Fig. 9.8, where all biasing resistors have been omitted from the drawing
for convenience. During normal operation the transistors should be in their active
region. For the best results, identical pairs of transistors are used and they will have
identical biasing components.
The BJT amplifier can be analyzed using the simple transistor model to get

−β Rc
Vout = (V1 − V2 ). (9.3)
rb

A real differential amplifier will not have components that are exactly matched and
will have an output that can be written

V2 + V1
Vout = Ad (V2 − V1 ) + Ac , (9.4)
2
where Ad is the “differential voltage gain” and Ac is the “common mode voltage gain.”
An ideal differential amplifier would have Ac = 0. One measure of the quality of a
differential amplifier is known as the “Common Mode Rejection Ratio” or CMRR,
ρ, where
 
 Ad 
ρ =  , (9.5)
Ac
9.2 Two-Transistor Configurations 183

and for a good differential amplifier ρ should be a large value (1). The CMRR is
usually expressed in decibels (dB)

CMMR in dB = 20 log10 (|Ad /Ac |). (9.6)

To achieve a large CMRR for the BJT example above, Re should be large. This can
be accomplished in practice by replacing Re with a constant current source (e.g.,
another transistor to produce a fixed current). The constant current source provides
the correct biasing, but looks like an infinite resistance for changes.
Ideally, the output of a differential amplifier, or any amplifier for that matter, should
not depend on (small) changes or transients from the power supply. A measure of
this is known as the power supply rejection ratio, or PSRR. It is defined in a manner
similar to the CMRR and is also usually expressed in dB. The appropriate ratio is
the change in the power supply voltage divided by the change in the output voltage,
for a fixed input.

9.2.5 Current Mirror

Another interesting application using two transistors is known as the “current mirror.”
The basic current mirror made with NPN transistors looks like the circuit of Fig. 9.9.
The circuit takes as an input the current I 1 and produces a copy, I 2 = I 1 , provided
the transistors are in their active regions and are closely matched in their properties.
When they are matched, the base currents through the two transistors will be identical,
and hence so will be the collector currents. Using identical transistors, any changes
in temperature, etc., will be the same for both and so the balance is maintained.
One use for such a circuit would be to create a controlled current source. For
example, the current I 1 is controlled using a known circuit, and that forces the current
I 2 to follow no matter what load is attached to it (provided the transistor stays in the
active region). Such circuits are also often found as part of the biasing circuitry for
more advanced integrated circuit amplifiers.

Fig. 9.9 The current mirror


uses two matched transistors
to produce the same current
in two separate paths
184 9 More on Amplifiers

9.2.6 Silicon Controlled Rectifiers (SCR) and Triacs

Another useful way to connect two transistors is shown in Fig. 9.10a. While some-
times used as a two-transistor device, this is accomplished in a single device known
as a silicon-controlled rectifier (SCR). An SCR consists of four layers with alternate
types of doping as shown in Fig. 9.10b. SCRs are sometimes also referred to as “thyris-
tors.”1 The schematic for an SCR is shown in Fig. 9.10c and the current–voltage
characteristics are illustrated in Fig. 9.10d.
The SCR has a double valued response,2 and so the solution depends on the history
of the device. If you start at zero (or negative) voltage, the current will stay small until
you reach a threshold voltage, at which point the current jumps to a large value. Once
you are at the larger current, the device looks like a resistor (with a low resistance)
until you get below a minimum holding voltage (or corresponding current), at which
point the device turns off.
The addition of a “gate” (or sometimes “trigger”) connection is used to control the
threshold voltage, and thus determines when the device turns on. To be completely
“on” typically requires a few tens of milliamps between the gate and the cathode.
That is, more or less, sufficient to “turn-on” the gate-to-cathode p–n junction.
The threshold voltage with no gate current can be 100 s of volts, and the minimum
holding voltage about 1 V. Usually the minimum holding current, rather than a
voltage, is specified and will be in the range of 1 to 100 mA for typical SCRs.
Note that the gate cannot be used to turn the device off. The device only turns off
when the anode to cathode voltage drops below the holding voltage (or current). As a
switch, this works for AC since a sinusoidal voltage will eventually go through zero
volts. DC power can be used if a separate switch is available to interrupt the current
and hence turn off the device—for example a “reset” button such as is found on an
alarm circuit.
Two SCRs connected in parallel, but in reversed directions, make a device known
as a triac. The schematic symbol for a triac is shown in Fig. 9.11. A triac produces a
symmetric output and is useful for AC power control. Triacs are used, for example,
in common household light dimmers and some motor speed controls. In those appli-
cations, the gate of the triac is used to turn on the signal somewhere in the middle of
the AC input cycle, then when the AC goes through zero the triac turns off. Hence,
the sinusoidal input power will be on for only a fraction of the total time each cycle.
If the gate turns on the signal just after each zero crossing, the AC will be on for
almost all of the cycle. On the other hand, if the gate turns on the signal just before
each zero crossing, the signal will turn off almost immediately, and little power gets
through.

1If the gate is omitted, the device may be referred to as a “four-layer diode” or a “Shockley diode”.
2The function is actually triple valued, however one of the solutions is always unstable and so the
observed response is double valued.
9.3 Connecting Amplifiers 185

Fig. 9.10 The two-transistor configuration in (a) can be constructed as a four-layer single device
(b), a silicon-controlled rectifier or SCR, that has the schematic shown in (c) and general
characteristics illustrated in (d)

Fig. 9.11 Two back-to-back


SCRs as a single device is
known as a triac, and is used
for switching AC signals

9.3 Connecting Amplifiers

In general, linear amplifiers such as those discussed so far (which are referred to as
“class A amplifiers”), no matter how constructed, can be modeled using Thevenin
equivalents. This is illustrated in Fig. 9.12. In that figure, Av is the “open circuit
voltage gain,” and Z i and Z o are the input and output impedances respectively. All
of the values may, of course, be frequency dependent. Also shown is a signal source,
with its equivalent output impedance, and the output of the string of amplifiers is
connected and a resistive load, RL . It is not hard to find the following relationships

Z 1i
V1i = Vs
Z s + Z 1i
Z 2i
V2i = Av1 V1o
Z 1o + Z 2i
RL
VL = Av2 V2i , (9.7)
Z 2o + R L

which describe the voltage amplification. Current and power amplification are treated
in a similar manner.
186 9 More on Amplifiers

Fig. 9.12 Thevenin’s theorem can be used to model the connection from a signal source, through
the amplifiers, and then to a final load, RL

9.3.1 Impedance Matching

Sometimes it is appropriate to maximize the power gain. For the example in Fig. 9.12,
the power supplied to the load,
 
Av2 V2i
PL = VL I L = RL , (9.8)
Z 2o + R L

for fixed gain Av2 will be a maximum when Z 2o = RL , a condition that is known as
an “impedance match.” To show this, find Z 2o that satisfies

d PL
= 0. (9.9)
d Z 2o

Likewise, the maximum power delivered to amplifier 2 occurs when Z 1o = Z 2i , and


the maximum power is delivered to amplifier 1 when Z s = Z 1i .
Impedance matching is also useful when the interconnecting cable lengths become
long. With longer cables, signals can reflect back to the source causing a length-
dependent impedance transformation (see Chap. 4). To avoid these reflections, the
characteristic impedance of the cable should match the input impedance of the
following amplifier. If maximum power transfer is also desired, then both of those
impedances should also match the output impedance of the original source.

9.4 Problems

1. Consider the circuit shown in Fig. 9.P1. Assume the transistors are in their active
region and have the same current gain β, and assume that all capacitors have
a small enough impedance for AC signals so that they can be considered short
circuits for those AC signals. Draw a carefully labeled circuit diagram appro-
priate for analyzing the small-signal AC behavior of this circuit. Such a diagram
includes replacing the transistors with the appropriate transistor model for the
9.4 Problems 187

active region, simplifying the diagram where possible, labeling currents, etc. It
is not necessary to do the analysis.

Fig. 9.P1 Problem 1

2. For the circuit of Fig. 9.P2, and assuming an ideal voltage source, prove that the
maximum power is delivered to RL when RL = Ro . In contrast, what value of R0
gives the maximum current through, and what value gives the maximum voltage
across, the load, RL ?

Fig. 9.P2 Problem 2

3. Derive the differential voltage gain for the two-transistor BJT differential ampli-
fier (i.e., derive Eq. 9.3). Assume the biasing has already been done, that the
transistors are in their active region, and that the transistors are exactly matched.
4. Use the FET transistor model to derive the differential voltage gain for the FET
differential amplifier. Assume any necessary biasing has already been done,
that the transistors are in their active region, and that the transistors are exactly
matched.
5. A differential amplifier with a differential voltage gain of 100 has an output of
0 V when the two inputs are both 0 V. If the amplifier has a CMRR of 60 dB,
what is the magnitude of the output when both inputs are 1 V?
Chapter 10
The Ideal Op-Amp

An operational amplifier (op-amp) is a special type of differential amplifier that has


some very useful properties. Op-amps are available as inexpensive self-contained
integrated circuits and the detail of what goes on inside the device is rarely of concern.
In this chapter, the ideal op-amp, treated as a device, is introduced and many useful
circuits based on the ideal op-amp with feedback are presented. After that, some
discussion of properties of a real op-amp are presented.

10.1 Ideal Op-Amp Properties

An ideal op-amp is a differential amplifier with the following additional properties:


• Input impedance infinite (no current enters the inputs)
• Output impedance zero (any current at output, as needed)
• Differential voltage gain is infinite
• CMRR infinite (i.e., common mode gain is zero)
• Frequency range infinite
• No offset voltages
• Characteristics are independent of power supply (PSRR infinite).
An op-amp as a device is shown symbolically using a triangle, as shown in
Fig. 10.1, in the same way as is any differential amplifier. One input signal is ampli-
fied with a minus sign, the “inverting input,” the other with a plus sign, the so-called
“non-inverting input.” The power supply connections to the op-amp are shown in
Fig. 10.1, however they are often omitted from a schematic diagram, or shown sepa-
rately from the op-amp, unless there is a good reason to include them. Of course, the
op-amp will not work without power, so those connections must be included during
circuit construction. Op-amps will usually have two power connections. For many
amplifier applications, one supply connection is usually a positive voltage and the

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 189
B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_10
190 10 The Ideal Op-Amp

Fig. 10.1 The op-amp


schematic, including power
connections

other negative.1 ,2 Often those two supplies are equal in magnitude, but there certainly
are exceptions. Since the device requires external power, an op-amp is referred to as
an “active device.”
Note that op-amps rarely have a ground connection—only the potential difference
between the two power connections is important for op-amp operation. Many of
the circuits that follow include a ground connection external to the op-amp. The
relationship between that ground and the two op-amp power sources is important for
proper circuit operation. Some op-amps have additional connections, not discussed
here, to allow for adjustments to make the behavior closer to ideal.
Op-amps are often used with feedback. Feedback is an external connection
between the output and the input. If the output of an amplifier is connected back
to the input, directly or indirectly, with a net negative sign, this is referred to as
“negative feedback.” Linear applications of op-amps use negative feedback of one
sort or another. Without that feedback, the large (ideally infinite) gain would result
in theoretically unrealistic, and practically unusable, output signals.
Negative feedback uses a correction signal to maintain the output at a certain level.
Imagine two cars on a highway. Car A is traveling at speed, V. Car B is supposed
to match the speed of car A. If car B is going too fast, the error in its speed, /\V, is
positive. To correct for this, the speed needs to be changed by −/\V. On the other
hand, if car B is going too slowly, the error in its speed is negative, −/\V. Now to
correct for that, car B needs to change its speed by −(−/\V ) = +/\V. In each case,
the error is multiplied by a minus sign to get the correction. That minus sign is what
is referred to as negative feedback. Using positive feedback, if car B is going too
fast, it will speed up even more, and it will never match speeds with car A.
When an ideal op-amp is in a circuit and has negative feedback, the output of
the op-amp will adjust, as best as it can, so that the two inputs to the op-amp are
equal. Many op-amp circuits can be solved using this special “device property.” Why
this technique works is demonstrated at the end of this chapter. First, a number of
useful example circuits are solved using this simplified device property. The second
important property used for these solutions, and important to remember, is that no
current enters (or leaves) either (ideal) op-amp input.

1 For general purpose op-amps, ±3 V to ±15 V would be typical. The appropriate range of values

for a specific op-amp will be found on the manufacturer’s datasheet.


2 A common beginner’s mistake during construction is to confuse the non-inverting (+) and inverting

(−) inputs with the positive (+) and negative (−) power supply connections.
10.2 Linear Op-Amp Circuits 191

10.2 Linear Op-Amp Circuits

Op-amps have utility both as linear amplifiers as well as for some non-linear appli-
cations, including digital circuitry. This section deals with linear applications. When
the op-amp is used with negative feedback, the circuit analysis can usually be simpli-
fied by assuming an ideal op-amp and, as mentioned above, that due to the negative
feedback, the output will adjust (if it can) to force the inverting and non-inverting
inputs to be equal. This method is shown by example in what follows.

10.2.1 Example 1—Buffer

Assuming an ideal op-amp, the relationship between the input and output voltages
for the circuit in Fig. 10.2 is easy to determine. First, write down the voltages at the
inverting (“−”) and non-inverting (“+”) inputs in terms of the voltages and any other
components shown. Use the fact that no current enters the op-amp inputs (an ideal
device rule). In this simplest case these input voltages are easily found to be

V+ = Vin , V− = Vout . (10.1)

There is negative feedback (due to the direct connection between the output and V − )
so we set V + = V − (the simplified “device rule” for the ideal op-amp when there is
negative feedback) and so clearly,

Vin = Vout . (10.2)

This circuit has a voltage gain of 1, but an extremely large current gain. The
current into the non-inverting input is extremely small and the output current can be
many orders of magnitude larger. The circuit is referred to as a buffer. It separates
the source of the input voltage from any load that may follow so that virtually no
power from the source is required. All the power is supplied by the op-amp power
supplies (not shown in the diagram).

Fig. 10.2 Example 1,


op-amp buffer
192 10 The Ideal Op-Amp

10.2.2 Example 2—Inverting Amplifier

In some circuits, it is beneficial to look at the current through the feedback component.
Consider the circuit of Fig. 10.3. The voltage at the non-inverting input is obviously
0 V (ground). Since there is negative feedback, the non-inverting (+) and inverting
(−) inputs are taken to be equal (i.e., V + = V − = 0 V). Then it is easy to see that I in
= V in /R1 . Since that current does not enter the op-amp input, it must go around (as
shown) through R2 . Hence, V out can be determined using the voltage drop from the
input,

R2 R2
Vout = V− − Iin R2 = 0V − Vin = − Vin . (10.3)
R1 R1

The voltage gain of the circuit is −R2 /R1 . Note that for this configuration, the gain
is completely under the control of the circuit designer. The input impedance of the
circuit, that is, the impedance “seen” by the input source, will be R1 . Be careful
not to confuse the input impedance of the op-amp (ideally infinite) with the input
impedance of the circuit, which includes the resistors.
Op-amp circuits with negative feedback and where the non-inverting input is
connected to ground (i.e., 0 V) will have a “virtual ground” at the inverting input.
The virtual ground will be 0 V, however unlike a real ground that can source and sink
current, the current here needs somewhere else to go.
In summary, the two basic strategies that are used to solve ideal op-amp problems
with negative feedback are:
• Compute the inverting and non-inverting input voltages in terms of the input
and output voltages and any additional components, then set the inverting and
non-inverting inputs voltages equal to each other and solve.
• Find the non-inverting input voltage, then set the inverting and non-inverting input
voltages equal. Now look at the voltage drop due to the current that goes through
the negative feedback component(s) and where that current came from and where
it can go.
Now look at some more examples that use these strategies.

Fig. 10.3 Example 2, R2


inverting amplifier
Iin
R1
Vin
(0V) Vout
Iin
10.2 Linear Op-Amp Circuits 193

10.2.3 Example 3—Non-inverting Amplifier

Figure 10.4 shows a non-inverting amplifier. It is obvious that V + = V in . Using the


voltage divider equation, valid since no current enters the op-amp inputs,

R1
V− = Vout . (10.4)
R1 + R2

There is negative feedback due to the connection between the output and V − , so V +
= V − , and

R1 R1 + R2
V+ = Vout → Vout = Vin . (10.5)
R1 + R2 R1

The input impedance for the non-inverting configuration is determined by the


particular op-amp used and will be very large.

10.2.4 Example 4—Difference Amplifier

Figure 10.5 combines the circuits of Examples 2 and 3 and uses two input voltages.
The voltage divider equation can be used for each of the op-amp inputs to get

Fig. 10.4 Example 3,


non-inverting amplifier

Fig. 10.5 Example 4,


difference amplifier
194 10 The Ideal Op-Amp

R4 R1 R2
V+ = VB , V− = Vout + VA . (10.6)
R3 + R4 R1 + R2 R1 + R2

Since there is feedback from the output back to V − , set V + = V − , and so


( )
R1 + R2 R4 R2
Vout = VB − VA . (10.7)
R1 R3 + R4 R1 + R2

For the special case where R1 = R3 and R2 = R4 ,

R2
Vout = (VB − V A ), (10.8)
R1

and this is a simple difference amplifier.

10.2.5 Example 5—Summing Amplifier

The circuit of Fig. 10.6 also has two inputs. Since the voltage at the non-inverting
input is fixed at zero (ground) and there is negative feedback through R3 , there is a
virtual ground at the inverting input. Hence, the two inputs provide a total current of

VA VB
I = I1 + I2 = + , (10.9)
R1 R2

which has nowhere to go except around the op-amp. Hence, starting at zero volts at
the inverting input and computing the drop across the feedback resistor, R3 ,
( )
VA VB
Vout = −R3 + . (10.10)
R1 R2

For the special case where R1 = R2 ,

Fig. 10.6 Example 5,


summing amplifier
10.2 Linear Op-Amp Circuits 195

R3
Vout = − (V A + VB ). (10.11)
R1

This configuration adds two signals together. If the two input resistors are not
equal, the circuit provides a weighted sum. This example was shown for two inputs,
but this idea can be extended to any number of inputs, each with its own weighting
resistor. If the minus sign on the output is of concern, the addition of an inverting
amplifier (Example 2) can be used to remove it.

10.2.6 Example 6—Integrator

Consider the circuit in Fig. 10.7. Since the non-inverting input is connected to ground,
there is a virtual ground at the inverting input provided that the capacitor gives
negative feedback—that is, since V + = 0, the feedback will keep V − = 0 as well.
Hence the current through R is I = V in /R. The solution here will use the same strategy
as Example 2 above, but with the additional complications due to the capacitor.
The current, I, does not enter the op-amp and so it must go to the capacitor. The
voltage across the capacitor will then grow with time as the charge builds up. The
voltage across the capacitor, V c , is given by

{t
Q I
Vc (t) = Vc (0) + = Vc (0) + dt, (10.12)
C C
0

and since there is a virtual ground (0 V) on the left side of the capacitor, and the
output on the other, it must be that V out = −V c .
Putting in the expression for the current, the output is related to the input by

{t
1
Vout = −Vc (0) − Vin (t)dt, (10.13)
RC
0

Fig. 10.7 Example 6,


integrating amplifier
196 10 The Ideal Op-Amp

and so this circuit provides an output linearly related to the integral of the input. It is
an “integrator.” If a provision for discharging the capacitor at t = 0 is included (for
example, a push-button switch across the capacitor) then the output is proportional
to the definite integral of the input starting at t = 0.
If the input is a sinusoid with amplitude V i at angular frequency ω, the circuit
could instead be analyzed using complex impedances. The type of analysis that is
appropriate will depend on the application.
Using complex impedances (and the fact no current enters the ideal op-amp)
proceed as in Example 2. The AC input current through R is V in /R and hence the
voltage drop from the inverting input to the output gives

Vin 1 i
Vout = − = Vin , (10.14)
R i ωC ω RC
a type of “low-pass filter” that has a large gain at small frequencies and a small gain
at larger frequencies. Remember that the “i” in the numerator is a 90° phase shift (a
sine becomes a cosine, a cosine becomes minus sine), so aside from a minus sign,
this is exactly what happens when you integrate a sinusoidal signal.
Any offset voltages on the inputs to the integrator are particularly annoying, since
the integral of a constant will simply grow linearly with time until the op-amp is out
of range. Such offsets may need to be addressed so that their effects are negligible.
One interesting application of the integrator is the circuit of Fig. 10.8. Four of
the resistors are taken to have the same resistance for simplicity, though that is not
necessary. The output, V out , is proportional to the integral of V 2 , which in turn is
proportional to the integral of V 1 . Hence, setting V 1 to

d2V
V1 = (RC)2 , (10.15)
dt 2
then
dV
V2 = −(RC) and Vout = V . (10.16)
dt
The inputs to the summing op-amp on the left will be

Fig. 10.8 A use of integrator circuits to model the driven harmonic oscillator
10.2 Linear Op-Amp Circuits 197

R2 R1
V+ = Vin + V2 V− = (Vout + V2 )/2. (10.17)
R1 + R2 R1 + R2

Since there is negative feedback, these inputs are equal. The resulting equation can
be written
( ) ( )
d2V 1 2R1 dV 1 1 2R2
+ + V = Vin , (10.18)
dt 2 RC R1 + R2 dt (RC)2 (RC)2 R1 + R2

which is same as the differential equation for a driven, damped harmonic oscillator
with natural frequency ω = 1/(RC). In electronics applications, this circuit is often
referred to as a “state variable filter.”3
It is straightforward to show that if the resistor and capacitor in Fig. 10.7 are
swapped, the output is, aside from some scale factors and a minus sign, the derivative
of the input.

10.2.7 Example 7—Low-Pass Filter

Figure 10.9 illustrates one of several circuits that can be used as a low-pass filter. That
is, lower frequencies are passed through while higher frequencies are blocked. When
circuits start to get more complicated, it is often useful to label some intermediate
voltages or currents to facilitate a solution. In this case, a label for the node voltage
V 1 has been added to the diagram.
There is a direct connection from the output to the inverting input and so there is
negative feedback and V out = V − = V + . Since no current enters the non-inverting
input, the voltage divider equation can be used to get

1/i ωC V1
V+ = V1 = = Vout , (10.19)
R1 + 1/i ωC 1 + i ω R2 C 1
or

V1 = Vout (1 + i ω R2 C1 ), so V1 − Vout = i ω R2 C1 Vout . (10.20)

Using KCL and Ohm’s law,

3 Taking outputs from V 1 , V 2 , and V out results in a high-pass, band-pass, and low-pass filters
respectively. The center frequency, gain, and damping terms can be adjusted independently.
198 10 The Ideal Op-Amp

Fig. 10.9 Example 7,


low-pass filter

I1 = I2 + I3
V1 − V+ V1 − Vout
I2 = = = i ωC1 Vout
R2 R2 (10.21)
V1 − Vout i ω R2 C1 Vout
I3 = = = −ω2 R2 C1 C2 Vout ,
1/i ωC2 1/i ωC2

and so

Vin = I1 R1 + V1 = (I2 + I3 )R1 + Vout (1 + i ω R2 C1 )


( ) (10.22)
= Vout 1 + i ω(R1 + R2 )C1 − ω2 R1 R2 C1 C2 .

Then
1
Vout = Vin (10.23)
1− ω2 R 1 R2 C 1 C 2 + i ωC 1 (R1 + R2 )

The low-pass behavior can be seen by considering limits. In the limit that the
frequency is very low (ω → 0) the denominator becomes 1 and V out = V in . When the
frequency is very high (ω → ∞) the denominator becomes very large in magnitude
and so V out → 0.
For comparison, solve the passive circuit of Fig. 10.10 using the voltage divider
equation to get

Fig. 10.10 Equivalent


circuit for Fig. 10.8
10.2 Linear Op-Amp Circuits 199

1
Vout = Vin . (10.24)
1− ω2 LC + i ω RC

This is of the same form as (10.23) and will have the same behavior if

LC = R1 R2 C1 C2 and RC = (R1 + R2 )C1 . (10.25)

Hence, Figs. 10.9 and 10.10 are, in a sense, equivalent circuits.4 Inductors are
notoriously non-ideal and tend to be large, heavy, and expensive compared to resistors
and capacitors, especially at lower frequencies. The circuit in Fig. 10.9 achieves the
same result without an inductor, and can, in fact, act more ideally than would real
components used with the circuit of Fig. 10.10. The so-called “cut-off” frequency
for the filter is where
√ the output is reduced by 3 dB, or in this case, the output voltage
is reduced by 1/ 2 from its low frequency value.
Swapping the capacitors and resistors in this circuit changes the response to that
of a high-pass filter. A slightly different arrangement can produce a band-pass filter.
These filters are generally referred to as “second-order active filters” and there are
several different configurations that can be used for each of them.

10.2.8 Example 8—Instrumentation Amplifier

It is often necessary to determine a small voltage difference, perhaps from a sensor


or the output from a bridge circuit (see Chap. 2). The difference amplifier above
(Example 4) can sometimes be useful; however, such an amplifier does not treat
the two input signals identically and may have an input impedance that is too small
to allow for an ideal voltage measurement. A much more symmetric and sensitive
treatment can be achieved using a so-called instrumentation amplifier, such as is
shown in Fig. 10.11. Note that the input impedance seen by both sources is Rin , which
is under the designer’s control and is often chosen to be relatively large (~1 Mu).
This is a three-op-amp design. Labels for the intermediate voltages V A' and VB' are
added to the diagram for convenience. The final op-amp (U3 ) is configured identically
to that of the difference amplifier of Example 4, and so5

R3 ( ' )
Vout = VB − V A' . (10.26)
R2

Both of the input op-amps (U1 and U2 ) clearly have negative feedback and thus
the voltage across Rg is V A − V B . No current goes into the inverting input of the
op-amps so the current that goes through both of the resistors R1 and the resistor Rg

4 To make the equivalence complete, a buffer (e.g., see Example 1) should be added to the output
of the circuit in Fig. 10.10 so that any load present will not draw current from the filter.
5 It is common to label integrated circuits, such as op-amps, using a “U.” Sometimes the two-letter

non-standard label “IC” is used instead of “U.”


200 10 The Ideal Op-Amp

Fig. 10.11 Example 8,


instrumentation amplifier

must be the same. Using Ohm’s law and equating the currents

V A' − VB' V A − VB
I = = , (10.27)
2R1 + Rg Rg

so
2R1 − Rg
V A' − VB' = (V A − VB ). (10.28)
Rg

Putting (10.28) into (10.26),


( )
R3 2R1
Vout = 1+ (VB − V A ). (10.29)
R2 Rg

High-performance instrumentation amplifiers can be purchased relatively inex-


pensively as a single pre-packaged device with laser-trimmed matched resistors.
Often there will be provisions to adjust the gain resistor (Rg ), that need not be
matched to any others, and possibly some other (usually optional) connections to
make the behavior even more ideal.

10.2.9 Example 9—A Capacitive Sensor for Smaller Values


of Capacitance

An op-amp circuit using a fixed known value of capacitance, C, can be used to


measure an unknown capacitance C x , using the circuit in Fig. 10.12a, where V (t) is
a known sinusoidal input voltage at angular frequency ω.
The resistance R is chosen to be a large value compared to 1/ωC and is only
necessary to provide a DC path to ground on the input, often necessary for a real
op-amp. It is ignored in the following analysis, which assumes an ideal op-amp.
10.2 Linear Op-Amp Circuits 201

Fig. 10.12 Example 9, (a) a circuit to measure a small capacitance, C x . In (b) the circuit is used
to detect a nearby conducting object, in this case a finger

Due to feedback, there is a virtual ground at the inverting input. The current into
the inverting input is

Iin = iωC V , (10.30)

which, due to the high input impedance of the amplifier, must go through C x . Hence,

Iin C
Vout = − = − V. (10.31)
i ωC x Cx

Since V (t) can be made very small, C x can be significantly smaller than C and the
output can still be easily measured. That measured value is used to determine C x .
This idea has been used to make a sensitive distance detector6 where the distance
between the plates of C x is measured. Alternatively, the capacitor configuration in
Fig. 10.12b has been used to detect fingers.7 In that case, the third capacitor “plate”
is the portion of the tip of a finger just above the sensor. The size of the such sensing
plates can be very small. The distance to the nearby conducting skin determines C x .
Many touch-screen and touch-pad devices use an array of capacitive sensors. The
fixed capacitor plates can be made from a layer of metal so thin that it is transparent.
Since such devices rely on the electrical conductivity of fingers to work, if a pen
or pencil is used instead of a finger, or even if the finger used is particularly dry,
capacitive touch devices will not respond.

10.2.10 Example 10—Negative Resistor

Consider the circuit in Fig. 10.13. This circuit differs from the previous exam-
ples in that it has an input but no output. This circuit demonstrates “impedance
transformation” using an op-amp.

6See, for example, Pigage et al. [1].


7For a similar application to detect fingerprints, see Tartagni and Guerrieri [2]. who use a charge-
based method rather than the AC impedance method described here.
202 10 The Ideal Op-Amp

Fig. 10.13 Example 10, a


negative resistor to ground

Assuming negative feedback is present so the inputs to the op-amp must be equal,
it is straightforward to write down

V ' = V − I R1
V = V ''
R2 R2
V '' = V ' = (V − I R1 ) (10.32)
R1 + R2 R1 + R2
( )
R2 R1 R2
V 1− = −I ,
R1 + R2 R1 + R2

which is easily solved to get V = −IR2 . That is, this circuit acts just like a resistor to
ground but with a negative resistance. The value of R1 does not show up in the final
equation and is determined by the current and voltage limitations of the device. That
is, R1 needs to be a value that keeps the device in a region of operation that provides
negative feedback, but otherwise its value is not important.
Note that in the circuit and analysis above, any impedance Z 2 could have been
used in place of R2 without changing the basic result. As a specific example, if a
capacitor were used in place of R2 , the circuit adds a minus sign to the capacitive
impedance, making the circuit act like an inductor to ground.

10.2.11 Example 11—Constant Current Source

There are several configurations that can be used for a constant current source. The so-
called Howland configuration uses a combination of positive and negative feedback.
Such a circuit could also be called a voltage-to-current converter. The circuit diagram
in Fig. 10.14 is referred to as the “improved Howland configuration.”
Assuming an ideal op-amp where the negative feedback is strong enough to win
out over the positive feedback, the output current is found using KCL, KVL, and the
properties of the ideal op-amp with feedback. With feedback, V + = V − ≡ V, and so
10.2 Linear Op-Amp Circuits 203

Fig. 10.14 Example 11,


constant current source

Vr e f − I1 R1 = V
V = I2 (R2 + R3 ) (10.33)
Vr e f − 2I1 R1 = V ' .

Combining these,

V ' = −Vr e f + 2I2 (R2 + R3 ). (10.34)

Now,

V '' = I2 (R2 + 2R3 ) = 2I2 (R2 + R3 ) − I2 R2 , (10.35)

so

V '' − V ' = Vr e f − I2 R2
V '' − V ' Vr e f
= −I ' = − I2 = −Iout − I2 (10.36)
R2 R2
Vr e f
Iout = − .
R2

Since the output current is independent of the load, the same output would be
expected for any load, at least to the extent the op-amp stays within its usable range.
The values of R1 and R3 , which do not directly affect the output current, are chosen to
ensure the op-amp stays within that range. Most real op-amps can provide a maximum
current up to no more than about 10 or 20 mA. If needed, the output current can be
amplified (e.g., using a BJT) to get much higher values.
204 10 The Ideal Op-Amp

10.3 Other Op-Amp Circuits

There are useful circuits that use the op-amp as a linear amplifier where the circuit
itself is not linear. Several such circuits are illustrated here as examples.

10.3.1 Example 12—Non-linear Element in Feedback

Consider the circuit of Fig. 10.15 where the FET characteristics are shown in
Fig. 10.16. Assuming the current I is positive, there is a connection, i.e., feedback,
from the output to the input through the (n-channel) FET. This circuit can be solved
with a graphical method.
Since the non-inverting input is connected to ground (0 V), the voltage at the
inverting input should also be 0 V and so I = 7 V/1.5 ku = 4.67 mA. The input
impedance of the op-amp is very large so this current must go through the FET.
Since the FET gate and source are connected together, V gs = 0, and the solution

Fig. 10.15 Example 12, an


op-amp with a non-linear
feedback component, in this
case an FET

Fig. 10.16 FET


characteristics used to solve
the circuit of Fig. 10.15
10.3 Other Op-Amp Circuits 205

must be on the V gs = 0 curve of the transistor characteristics. Follow the curve corre-
sponding to V gs = 0 V until I ds = 4.67 mA. Now read off V ds . Thus, for this circuit
V out = −V ds = −1.8 V.
Note that if the input voltage were 15 V then I = 9.33 mA if negative feedback
is assumed. However, no solution exists for I = 9.33 mA. This, indirectly, shows
that the feedback cannot make the two inputs equal and the assumption of negative
feedback is invalid for this case.8 It is necessary to go back and rethink the problem.

10.3.2 Example 13—Ideal Diode

As far as the voltage is concerned, the circuit in Fig. 10.17 acts almost like an ideal
diode. There will be a connection between the output and inverting input only when
the diode is forward biased. When the diode is reverse biased, the feedback is broken
and the output of the amplifier will tend toward minus infinity. In practice the output
will end up at the maximum negative value the op-amp can provide (which is usually
close to the negative power supply voltage).
Assuming that there is negative feedback, the two inputs to the op-amp will be
equal, and hence V out = V in . It is clear, however, that the diode is only conducting
when V out > 0. The assumption of negative feedback is only valid when V in > 0.
When V in < 0, the output of the op-amp goes negative and the feedback is broken.
With the diode reversed biased, no current is supplied to the load (R) and so the
output voltage at the load, V out , is zero.
Hence, for the ideal op-amp, the output voltage is what you expect for an ideal
diode:

Vin > 0 → Vout = Vin


(10.37)
Vin < 0 → Vout = 0.

A more careful analysis shows that the effective diode turn-on voltage is not quite
reduced to zero, but is reduced by a factor comparable to the gain of the op-amp (e.g.,
a factor of 105 or more). Hence, instead of having a turn-on voltage of about 0.5 V,
the effective turn-on voltage is about 5 µV for a typical op-amp. That is negligible
compared to typical op-amp offset voltages (~1 mV) present for real op-amps. Note
also that this circuit only acts as an ideal diode for the input voltage. Any current
from the input source is blocked by the op-amp.
During the negative part of the input signal, the op-amp output goes as low as it
can. Since it can take a while for the op-amp to recover from that, this circuit may
only work well at lower frequencies. How low will depend on the particular op-amp.
Should that frequency dependence be an issue, there are similar circuits that use
strategies that provide an alternate connection for the output when the diode is off,

8 It is always good to check assumptions, even if a solution does exist.


206 10 The Ideal Op-Amp

Fig. 10.17 Example 13, an


ideal diode circuit

and that alternate path keeps the output from the going to the extremes. One of these
is schemes is included in the next example.

10.3.3 Example 14—Peak Follower

Consider the circuit of Fig. 10.18. The second op-amp, U2 , is a buffer (see Example
1) and hence the output will be equal to the voltage across the capacitor, V C . The
output from the first op-amp (U1 ) is labeled V 1 .
If V 1 is greater than V C then diode D1 is conducting and diode D2 , which sees
V C via the output of U2 , is “off.” In that case, U1 gets negative feedback from the
output of U2 , equivalent to the feedback in the previous example. The output of U1
follows the input and the capacitor follows along.
If V 1 is less than V C , then diode D1 is “off” and diode D2 conducts. With diode
D2 conducting, U1 has negative feedback directly from its own output and it forms a
buffer, so that V 1 = V in . Since that signal is blocked by D1 , it is of no real consequence
except that U1 is prevented from going to its extremes. So if V in is smaller than V C ,
the capacitor is disconnected from the input. When disconnected, the capacitor will
eventually discharge through R1 , there being no other place for the current to go.
If resistor R1 is absent, the capacitor will charge up to the maximum value of V in
and just sit there, storing the peak value of the signal. Including that resistor allows
the capacitor to “reset” with time. Alternatively, a (transistor) switch could be used
to discharge the capacitor once the stored peak value is no longer required.
How rapidly the circuit will recover from a peak in the input depends on the R1 C
time constant. Figure 10.19 illustrates the behavior for a very large and a moderate

Fig. 10.18 Example 14, a peak follower


10.3 Other Op-Amp Circuits 207

Fig. 10.19 A short audio


signal (solid line) sent to the
peak follower saves the peak
if a large time constant is
used (dashed) or follows the
“envelope” of the signal if a
moderate time constant is
used (dotted)

time constant. The latter might be used to try to follow the envelope of the input. If
the time constant is made very short, for example if the capacitor is simply removed,
the output is the same as that of the ideal diode circuit (Example 13).

10.3.4 Example 15—Log Amplifier

If the input voltage, V in , shown in Fig. 10.20, is positive, then the diode will be
conducting. With the diode “on” there is feedback—the diode acts like a wire for
changes in the voltage or current—and hence a virtual ground at the inverting input.
The input current is then I = V in /R. The output voltage is then determined from the
voltage drop across the diode, V d . That is, V out = −V d . Using the analytic model
for the diode (6.1), the input and output voltages are related by

( ( ) )
Vin Vout
I = = Id = I0 exp − −1 (10.38)
R ηVT

Recall that for a semiconductor diode, I 0 is very small, so that if I d is any appre-
ciable value (say, greater than about 1 µA), then the exponential must be very large

Fig. 10.20 Example 15, a


log amplifier
208 10 The Ideal Op-Amp

compared to 1 and the “−1” can be neglected in comparison. The output in terms of
the input can be found by multiplying both sides by R, taking the log of both sides,
and then using properties of the logarithm. That is,

log(Vin ) = log(R I0 exp(−Vout /ηVT )) = −Vout /ηVT + log(R I0 )


(10.39)
Vout = −ηVT log(Vin /R I0 ).

Aside from some scale factors and a minus sign, the output is the logarithm of the
input.
Note that if the input voltage is negative, the diode does not conduct and the
feedback is broken. The op-amp will go to its largest possible output voltage. Since
there are no (real) logarithms for negative numbers, such behavior should not be
surprising for negative inputs.
A logarithm circuit is useful for indicating signal levels in decibels (dB), since,
after all, dBs are a logarithm with a scale factor. Common signals with an amplitude
that is convenient to express in dB include those that arise from the detection or
transmission of sound, light, and radio signals.9
If the positions of the diode and resistor are swapped, the circuit becomes an
“anti-log” or exponential circuit. The combination of log and anti-log circuits, along
with sum and/or difference circuits, can, in principle, be used to multiply or divide
two values. It is difficult to get a precision multiplication from such circuits due to
imperfections in the diode model and the temperature dependence, particularly of
VT.

10.3.5 Example 16—Absolute Value Circuit

Figure 10.21 illustrates a cousin to the ideal diode circuit. Here op-amp U2 is
configured as a summing amplifier (see Example 5) and so
( )
Vout = − Vin + 2V '' . (10.40)

If V ' > V in , then diode D2 conducts and there is negative feedback and a virtual
ground at op-amp U 1 . With a virtual ground at both op-amps and D1 off, it must
be that V '' = 0. If V ' < V in , then D2 is off and U1 gets negative feedback through a
resistance R, forcing V '' = −V in . There is feedback in either case. The former occurs
when V in < 0 and the latter when V in > 0. Thus,

If Vin < 0, Vout = −(Vin )


(10.41)
If Vin > 0, Vout = −(Vin − 2Vin ) = Vin ,

9One place to observe such measurements is the “VU meter” that may be on the output of an audio
amplifier or on an audio sound board.
10.4 More Power 209

Fig. 10.21 Example 16, an absolute value circuit

or, equivalently, V out = |V in |. Because of this, this circuit is sometimes referred to as a


full-wave rectifier, however, as for the ideal diode (Example 13), no significant current
from the input passes to the output. To get good results with this circuit, the resistors
should be well-matched. To help achieve a good match with the other resistors, the
resistor with value “R/2” can be obtained using two equal valued resistors, R, in
parallel, so that all the resistors used have the same value and can be matched to
great precision, for example by measuring each in turn using the Wheatstone bridge
seen in Chap. 2.

10.4 More Power

Op-amps are low power devices used primarily to process or create signals. You
should not expect to get a significant amount of power from an op-amp. Op-amps
will not run motors, speakers, or light up bright lights by themselves. In simple cases
where more power is required transistors can be added in a manner similar to the
circuits in Fig. 10.22.
The first circuit (Fig. 10.22a) might be used to control a positive voltage across a
load such as a small motor or a modest heating element. The feedback will cause the
voltage across the load to match the input voltage, as long as the values stay within
the ranges allowed by the devices.
This second circuit (Fig. 10.22b) might be used for audio signals, or other signals
that are both positive and negative. Again, the feedback causes the output across the
load to follow the input voltage. The resistors shown, other than the load, help to keep
the current within acceptable ranges and will depend on the particular application.
For any such circuits, where feedback is taken from the output of the power
transistors, the tendency is that any attempts to achieve more power at the output will
likely reduce the maximum usable frequency.
210 10 The Ideal Op-Amp

Fig. 10.22 Two circuits showing the use of transistors to boost the output power from an op-amp

10.5 Less than Ideal Difference Amplifiers

10.5.1 Finite Input Resistance and Gain

The solution for a general differential amplifier is a bit more complicated to obtain
than it is for the ideal op-amp. The main purpose of this section is to illustrate how
you might do that and then to look at the limit as the amplifier becomes ideal. The
results found in that limit justify the simplified solution method used for all the
previous op-amp examples.
Consider the simple circuit shown in Fig. 10.23 that is constructed using a differ-
ential amplifier with finite differential voltage gain, A, and an (effective) internal
resistance between the inputs, Ri . In terms of the input voltage, V in , and unknown
output voltage, V out , the inputs to the differential amplifier can be found using simple
superposition.

V+ = 0
Ri ||R2 Ri ||R1 (10.42)
V− = Vin + Vout ,
Ri ||R2 + R1 Ri ||R1 + R2

Fig. 10.23 A differential


amplifier including an input
impedance between the two
inputs
10.5 Less than Ideal Difference Amplifiers 211

and these can be substituted back into the expression for V out :
( )
Ri ||R2 Ri ||R1
Vout = A(V+ − V− ) = −A Vin + Vout . (10.43)
Ri ||R2 + R1 Ri ||R1 + R2

Now solve this for the output voltage to get


Ri ||R2
Ri ||R2 +R1
Vout = Ri ||R1
Vin . (10.44)
1 + A Ri ||R 1 +R2

Now if the input resistance, Ri , is very large compared to both R1 and R2 , this
reduces to

−A R2R+R
2

Vout = 1
, (10.45)
1 + A R1R+R
1
2

and in the limit that A becomes very large this is

R2
Vout = − Vin , (10.46)
R1

as expected (e.g., see the inverting amplifier, Example 2 above). Now to justify
the simpler solution method used for ideal op-amps, put this solution back into the
expression for V − (with Ri large),
( )
R2 R1 R2 R2 R1
V− = Vin + Vout = Vin + − Vin = 0,
R1 + R2 R1 + R2 R1 + R2 R1 R1 + R2
(10.47)

and it can be seen that the feedback has forced V − = V + in this limit.
This problem can be addressed in a completely different way using Miller’s
Theorem (see Chap. 9). Using Miller’s theorem, the circuit of Fig. 10.23 is equivalent
to the circuit in Fig. 10.24.
If R2 << (1 + A)Ri , a condition easy to meet if both A and Ri are large, then

Fig. 10.24 An equivalent to


the circuit of Fig. 10.23
obtained from Miller’s
theorem
212 10 The Ideal Op-Amp

R2 /(1 + A)
V+ = 0 ; V− = Vin
R + R2 (1 + A)
(2 ) ( )
R2 /(1 + A) R2 A/(1 + A)
Vout = −AV− = −A Vin =− Vin
R1 + R2 (1 + A) R1 + R2 (1 + A)
(10.48)

and the previous result is obtained in the limit that A becomes very large. Referring
back to Fig. 10.24, note also that once again as A becomes large, V − approaches V +
due to the feedback.
Thus, a real differential amplifier with feedback will act like an ideal op-amp if
the amplifier gain, A, is very large compared to 1 and the amplifier input impedance
is large compared to the feedback impedance.
Available integrated circuit op-amps typically have an “open loop gain,” i.e., with
no feedback, between about 105 and 106 , which is certainly quite large compared
to 1. Op-amp voltage gain is so large it is often expressed in units of “volts per
millivolt,” or V/mV, so a voltage gain of 105 would be written as 100 V/mV.
Op-amp input impedances can range from about 106 to 1012 u. The latter is
comparable to the resistance of typical wire insulation, circuit board material, and
many other insulators—hence, if such high impedance values are of importance,
some care must be taken to avoid reducing that resistance through nearby insulating
materials. With op-amps, one crude, but simple method to do this is “dead bug
construction.” The op-amp case is glued to a mounting board with the connecting
leads pointed up in the air, thus avoiding any contact with mounting sockets, circuit
board material, or other insulators.

10.5.2 Finite Frequency Range

A general purpose, inexpensive integrated circuit op-amp can be expected to have a


unity gain bandwidth of about 1 MHz. That is, the buffer of Example 1 can be used
up to 1 MHz, the inverting or non-inverting amplifiers (Examples 2 and 3) with a
gain of 10 can be used to 100 kHz, and so on. Such op-amps are usually adequate for
signals in the audio range (0–20 kHz). With slightly more expense, op-amps with a
unity gain bandwidth into the 100s of MHz are available.
While it may be tempting to use an op-amp with a very high frequency response for
lower frequency operation, after all that would be more “ideal,” such use can lead to
trouble. Remember that the amplifier is still working at its highest frequencies even if
it is not being used at those high frequencies. At 100 MHz and above, even relatively
short wire lengths become very important (see Chap. 4) and can act as impedance
transformers. With a high frequency op-amp, the circuit design and layout need to
take the high-frequency response into account even if the input signals are much
lower in frequency. It is very easy make such circuits behave very poorly due to
10.6 Oscillations 213

a bad layout or to have these high-frequency amplifiers turn into high-frequency


oscillators10 due to unintended inappropriate feedback.

10.5.3 Small Signals and Drift

An amplifier’s offset voltages can be a nuisance when trying to amplify very small
signals (~1 mV). While the offset can be compensated at any given time, offsets tend
to drift over time, especially with changes in device temperature. There are various
schemes to deal with this issue and there is not time and space to go into all of them
here. One particularly simple scheme is to turn a relatively slowly varying input
signal into a time dependent one. Since the drift is slow and appears as a d.c. offset,
that offset can be removed with a capacitor. Such a “chopping” scheme is illustrated
Fig. 10.25, where the peak-to-peak amplitude of the output is proportional to the
low-level signal input. The electronically controlled switch would be implemented
with a timing circuit and transistor switches (available as integrated circuits). Raising
the frequency of the desired signal before the amplifier also has some advantages
for noise reduction since amplifiers tend to produce more electrical noise at lower
frequencies than at higher frequencies.

10.6 Oscillations

Any amplifier with positive feedback can oscillate if conditions are right. Consider
the non-inverting circuit of Fig. 10.4 (Example 3). Then

R1
V− = Vout ; V+ = Vin , and so
R1 + R2
(10.49)
R1
Vout = A(V+ − V− ) = AVin − A Vout ,
R1 + R2

or defining the fraction of the output that is used for feedback, β, as

R1
β=− , (10.50)
R1 + R2

where the minus sign indicates negative feedback is being used. Then

A
Vout = Vin . (10.51)
1 − βA

10 Such unintended oscillation is referred to as “parasitic oscillation.”


214 10 The Ideal Op-Amp

Fig. 10.25 One scheme to


reduce problems with
low-frequency drift is to
convert the input signal to a
higher frequency using an
on/off switch

In Example 3, β < 0 (so βA is negative), so all is as before.


Now consider what happens for the circuit of Fig. 10.4 if the two inputs are
swapped. That has the same effect as changing A into −A. Then if components are
chosen so that βA = 1, the denominator goes to zero in which case an output is
expected even in the limit that the input becomes negligibly small. This is known as
the Barkhausen criterion. That criterion is much more interesting when capacitors
and/or inductors are involved in the feedback. Then β can become a frequency
dependent complex value. If A = 1 at some non-zero frequency, an output signal at
that frequency can be expected even in the absence of an input. That is, the circuit
oscillates.
One example of an op-amp oscillator is the twin-T oscillator, shown in Fig. 10.26.
The circuit can be solved using KCL, the virtual ground, and the currents shown.
The result is

1 − i ω RC
Vout = Vin . (10.52)
1 − (ω RC)2

Fig. 10.26 The twin-T


oscillator made using an
op-amp
10.8 Problems 215

When ωRC = 1 the denominator goes to zero, and oscillation should be expected.11
The combined phase shifts from the capacitors have turned what looks to be negative
feedback into positive feedback.

10.7 The Transconductance Amplifier

A cousin to the op-amps discussed above is the so-called operational transconduc-


tance amplifier (OTA). Like the op-amp, there are two high impedance differential
inputs. However, the output is a current rather than a voltage. That is,

Iout = gm (V+ − V− ). (10.53)

The output voltage will depend on the load. There is often an additional separate
current input that can be used to control gm . Applications often take advantage of
that extra control to make voltage-controlled volume controls, voltage-controlled
filters, and similar circuits.

10.8 Problems

• (a) Show that for sinusoidal input signals, the circuit in Fig. 10.P1 gives an output
that has the same magnitude as the input, but a different phase, and that the resistor
R can be adjusted to get different phase shifts. (b) What is the maximum range
of the phase shifts that can be expected? Assume an ideal op-amp. This circuit is
sometimes called an “all-pass filter” since all frequency components pass through
with no change in amplitude.

Fig. 10.P1 Problem 1

11 In the language of Chap. 5, there are poles at s = ± i/(RC).


216 10 The Ideal Op-Amp

• For the circuit in Fig. 10.P2, how is the output related to the input for the case
where the switch is open and when it is closed?

Fig. 10.P2 Problem 2

• Compute the voltage gain (V out /V in ) for the circuit in Fig. 10.P3, assuming an
ideal op-amp.

Fig. 10.P3 Problem 3

• In the circuit of Fig. 10.P4, the non-linear circuit element “x” has the current–
voltage relationship Vx = bIx3 , where b = 1.5 V/(mA)2 is a constant. What is the
output voltage, V out ?

Fig. 10.P4 Problem 4


10.8 Problems 217

• A solar cell (or photovoltaic cell) is a two-lead device used to detect and extract
electric power from light. A small solar cell can be used as the detector for a
light meter. Such a cell connected to a short circuit will emit a current roughly
proportional to the number of incident photons that are within the cell’s detection
range. Design a circuit with a solar cell, an op-amp, and resistors (as required)
that provides an output voltage proportional to the number of incident photons.
• Assume ideal op-amps and derive the relationship between V out and the two inputs,
V A and V B for the circuit in Fig. 10.P6.

Fig. 10.P6 Problem 6

• Show that the op-amp circuit in Fig. 10.P7a behaves the same as the circuit in
Fig. 10.P7b. These circuits act as a band pass filter, allowing signals to pass
through only if their frequency is near to the filter’s center frequency. If the center
( √ )−1
frequency for the circuit in Fig. 10.P7b is f 0 = 2π LC , what is the center
frequency for the op-amp circuit?

Fig. 10.P7 Problem 7

• Derive V out /V in for the op-amp circuit in Fig. 10.P8. Assume an ideal op-amp.
• Assuming an ideal diode and op-amp, show that the output voltage of the circuit
in Fig. 10.P9 is the absolute value of the input voltage. How does the behavior
change for a real semiconductor diode?
218 10 The Ideal Op-Amp

Fig. 10.P8 Problem 8

Fig. 10.P9 Problem 9

• For the circuit of Fig. 10.P10 what is V out ? Assume ideal op-amps and that for
any input one of the diodes is conducting. Consider both positive and negative
V in . Why will it never be the case that both diodes are conducting?

Fig. 10.P10 Problem 10

• For the circuit of Fig. 10.P11, assume ideal op-amps with feedback and deduce
the currents I 1 and I 2 . (Hint: the answers are simple though perhaps somewhat
unusual).
10.8 Problems 219

Fig. 10.P11 Problem 11

• (Challenge Problem) Assume ideal op-amps and compute the input impedance
Rin = V in /I in and the current through R5 for the circuit shown in Fig. 10.P12.
(Hint: the answers are relatively simple mathematical relations).

Fig. 10.P12 Problem 12

• (Challenge Problem) Analyze the circuit shown in Fig. 10.P13 assuming ideal
op-amps with negative feedback to determine V 0 in terms of V 1 and V 2 . Why is
it ok to assume negative feedback is being used in this circuit even though the
upper op-amp has feedback applied only to the non-inverting input?
220 10 The Ideal Op-Amp

Fig. 10.P13 Problem 13

References

1. R.F. Pigage, A.B. Ryan, R.E. Smith, Method for Inspecting and Adjusting Cutter Blades, U.S.
Patent No. 3,641,431 (1968)
2. M. Tartagni, R. Guerrieri, A fingerprint sensor based on the feedback capacitive sensing scheme.
IEEE J. Solid-State Circ. 33, 133–142 (1998)
Chapter 11
Non-linear Uses of Op-Amps

In the previous chapter, the op-amp was used as a linear amplifier. A real op-amp
has limitations that can be exploited to make useful circuits in which the op-amp
is no longer linear. These applications use the very high gain of the op-amp along
with the limited output range to create new circuits. Of particular interest is use as a
comparator.

11.1 Limited Output Range

The output range of integrated circuit op-amps is limited, or “clipped,” near the level
of the power supplies. That is, if V+ is the positive supply and V− the negative,1 then
the output voltage, V out , is restricted to the range (V− + δ) < V out < (V+ − δ), where,
for general purpose op-amps, δ ≈ 1 V. There are slightly more expensive op-amps,
the so-called “rail-to-rail” op-amps, where δ can be as small as a few millivolts.
The inputs for most op-amps are also limited by the supply voltages. For general
purpose op-amps, the inputs are bounded by the supply voltages—they should be no
smaller than V− and no larger than V+ . For some more expensive op-amps, which
includes many of the “rail-to-rail” op-amps, the inputs can exceed the supplies by
as much as a volt without causing damage. The manufacturer’s datasheet for the
op-amp should be consulted for details.
Provided the op-amp extremes are not exceeded, the limiting behavior can be of
use for some interesting applications. Consider, for example, the circuit of Fig. 11.1,
assumed to be constructed using rail-to-rail op-amps.2 The “negative” supply is set

1 Power supply connections are often labeled using ±V cc , V cc and V ee , and/or with V dd and V ss .
V cc and V dd being the higher voltage of their pair. The letters in the subscripts originally referred
to the collector, emitter, drain, and source of the transistor(s) in the amplifier.
2 For example, using the LMC6482 dual op-amp, R = 10 k and +V = 5 V.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 221
B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_11
222 11 Non-linear Uses of Op-Amps

Fig. 11.1 Absolute value circuit using rail-to-rail op-amps

at ground (0 V). The first op-amp (U1 ) is wired as a buffer, but the output cannot
go negative. Hence U1 only amplifies when the input is positive. The output of the
first op-amp will look like that of a half-wave rectifier. When the input is positive,
the second op-amp (U2 ) will see V in from U1 at its non-inverting input. The negative
feedback will adjust so the inverting input is also equal to V in , and it does that by
adjusting the output voltage to V in . When the input is negative, U1 outputs 0 V, since
it is limited by the power supply, and that creates a virtual ground at the inverting
input of the second op-amp. The negative feedback on U2 adjusts the output to −V in
(a positive value) to achieve this. Thus, this combined circuit outputs the absolute
value of the input without using any diodes.
There may seem to be several more resistors in Fig. 11.1 than are really necessary.
For example, the two connected to the inputs of U1 and the resistor to the non-
inverting input of U2 could be wires and the circuit analysis would look the same.
These resistors will ensure that the voltage at the inputs to the (real) op-amps do
not go too far outside the supply range. Those “extra” resistors limit the possible
current that can occur. The values of those resistors are not too critical. The device
datasheet will provide values for the maximum allowable voltages and currents for
the particular device used.

11.2 The Op-Amp Comparator

The simple circuit in Fig. 11.2, an op-amp without feedback, creates a comparator.3
For this circuit, the difference between the two input voltages is greatly amplified.
For the ideal op-amp, the output would approach ± infinity. As mentioned above, for
a real op-amp the output is limited to values between the two supply voltages. Thus,
for the real op-amp, the two input voltages are compared and the output has one of
two values, the maximum and minimum that the op-amp can produce, depending on
which input is larger. For the real op-amp, this will be true except for a very small

3A “digital comparator” will compare two digitally stored numeric values. Be aware that the
word “digital” is sometimes omitted. Digital comparators are functionally quite different from the
comparators discussed here.
11.2 The Op-Amp Comparator 223

Fig. 11.2 Schematic for a


comparator—an op-amp
without feedback

range of values where the two inputs are extremely close together—closer than about
10 μV or even less. The comparator is a highly non-linear application for the op-amp.
In summary, V out is near the positive supply voltage, V+ , if V B > V A or the negative
supply voltage, V− , if V B < V A . In practice, making V A = V B to get an output ½-way
in between is virtually impossible for a real op-amp and that possibility is simply
ignored here.
The comparator is useful for many applications and several examples are shown
below. It should not be surprising that comparators are often part of the interface
between analog electronics and digital (on/off) electronics.
There are devices available that are designed especially to be used as comparators.
Some of the circuitry that is added to op-amps to make them behave more ideally
as amplifiers will compromise the behavior when used as a comparator. That addi-
tional circuitry can simply be omitted. For many applications, particularly at lower
speeds, an op-amp works just fine as a comparator. In what follows, no distinction is
made between op-amps used as comparators and these devices which are especially
designed to be used as comparators.
If a small amount of positive feedback is added to a comparator, the result is a
“comparator with hysteresis.” The input/output relation is double-valued and depends
on the previous history of the device, not just the instantaneous values of the inputs.
Such a circuit is illustrated in Fig. 11.3a. Here, V in is compared to the voltage at the
non-inverting input, V + , which is in turn given by

Rp Rr
V+ = Vr e f + Vout . (11.1)
Rr + R p Rr + R p

Fig. 11.3 A comparator with positive feedback (a), has a double valued output (b), with
hysteresis—the output depends on previous values of the inputs
224 11 Non-linear Uses of Op-Amps

Hence, V + depends on the present state of V out . And in turn, V out is near either the
positive or negative power supply voltage (V+ or V− ) depending on the previous state
of the input. Tracing the behavior starting with V in = V− , sweeping up to V+ and
then back down to V− will produce the double-valued behavior shown in Fig. 11.3b.
The extra threshold to switch, created by the positive feedback, is useful for noise
immunity. Comparators with hysteresis are available as a self-contained device for
certain applications.4
A number of applications using comparators follow. For many of them, hysteresis
could also be included, but is not shown.

11.2.1 Example 1—Low-Level Warning

Comparators can be used to provide a simple indication of a signal level. In the


simplest case, it might simply provide a warning when a signal, such as a battery
voltage, is too high or too low.
A simple indicator that provides a warning if a supply voltage has dropped below
a preset threshold is shown in Fig. 11.4. In this circuit, the diode on the right is an
LED that glows when there is a forward current of a few mA. The forward voltage
drop of the diode on the left, its turn-on voltage, is being used to provide a reference
voltage (about 0.5 V) which is relatively independent of the power supply voltage.
A Zener diode (with reverse bias) could be used instead. When the voltage at the
inverting input drops below the diode voltage, the LED turns on. So, for example,
if this circuit is used with a 9 V battery and a warning is desired when the battery
voltage drops below about 8.8 V, choose R1 ≈ 100 k and R2 so that

R2
VD = 8.8 V, (11.2)
R1 + R2

where V D is the (measured) diode reference voltage for your circuit.


A more complicated circuit may use several comparators to indicate a more contin-
uous range of values, such as for the audio level on some audio devices. Figure 11.5
shows such a circuit that will indicate three levels. The same principle can be used for
many more levels.5 The resistors on the right are included to limit the current through
the LEDs. They may not be necessary if the output current from the comparators
(op-amps) is internally limited at a safe level.

4 A comparator with hysteresis used for digital electronics is usually referred to as a “Schmitt
trigger.”
5 The LM3914 and LM3915 integrated circuits include 10 such levels built into a single integrated

circuit device.
11.2 The Op-Amp Comparator 225

Fig. 11.4 Example 1—a


simple low-voltage warning
circuit using a comparator

Fig. 11.5 A three-level


voltage indicating circuit.
Such a circuit can be
expanded to many more
levels

11.2.2 Example 2—Pulse Generator

The circuit of Fig. 11.6 can be used to make a voltage pulse. The pushbutton switch
here is normally open (N.O.), meaning that there is no connection until the switch is
pushed. When the switch is pressed (closed), the capacitor is rapidly charged to +V,
which will exceed the value at the inverting input. Once the switch is released, the
capacitor will gradually discharge through resistor R0 . The output of the comparator
will remain high (i.e., near V+ ) until the capacitor voltage falls below the voltage at
the inverting input. Then the output goes low (i.e., near V− ).

Fig. 11.6 Example 2—a single pulse generator. When the pushbutton is pressed and released, a
square pulse is generated
226 11 Non-linear Uses of Op-Amps

A typical mechanical switch makes and breaks the connection between two
conductors to perform the switching action. As it does so, the connection actually
makes and breaks many times over a short time period (typically a few milliseconds),
a phenomenon known as switch bounce. For many applications that is not a problem,
but for high-speed digital applications, each make and break is treated as a separate
event, as if the switch were repeatedly pushed and released. This pulse generator can
eliminate that effect, and when it is used for that purpose, it is referred to as a switch
de-bouncer.

11.2.3 Example 3—Simple Oscillator

Figure 11.7 shows a comparator with hysteresis with a capacitor and resistor added
to supply the inverting input voltage. This creates a simple oscillator.
Start with the capacitor discharged and the op-amp output at either of its two
possible values (“high” or “low”). Suppose it is high. Then the non-inverting input
will see a positive voltage and the capacitor will begin to charge upward. Once the
capacitor reaches the value at the non-inverting input, the op-amp output will switch
and go low. Then the non-inverting input will see a lower value and the capacitor will
start to discharge toward that lower value. Once the capacitor reaches the lower value,
the op-amp output once again goes high, and the process repeats. It is somewhat like
a game of keep-away, where the capacitor is trying to get to the target value at the
non-inverting input, but each time as it just gets there, the target changes.
There are two outputs indicated in Fig. 11.7. The output V out1 will be a square
wave. The voltage across the capacitor, V out2 , will be that of the capacitor charging
and discharging. If the hysteresis loop is not too large, V out2 will approximate a
triangle wave.

Fig. 11.7 Example 3—a simple oscillator made using a comparator with hysteresis that uses a
delayed version of its own output as one of the inputs
11.2 The Op-Amp Comparator 227

Fig. 11.8 Example 4—a voting circuit

11.2.4 Example 4—A Voting Circuit

The circuit in Fig. 11.8 will function to determine the outcome of a voting process.
The first op-amp sums the voltages from all the inputs. In this case a closed switch
is a “no” vote and an open switch is a “yes” vote. The output of the first op-amp
varies from 0 V (all “no”) to −V ref (all “yes”). The comparator checks to see if the
result is more or less than −V ref /2. The output, V out , will be near the positive supply
voltage if the majority say “yes” and near the negative supply voltage if the majority
say “no.”
Since real resistor values are not particularly accurate and may depend somewhat
on temperature, such a circuit could only be used for a relatively small number of
“voters.” It would also be best to use it only for an odd number of voters since the
results are unpredictable in the event of a tie. Such a “voting” procedure may not
necessarily be part of a democratic process. For example, it might be desirable to
have an alarm on an airlock that goes off if any 2 of 3 doors are open. In that case
this circuit can be used with n = 3 and switches appropriately wired to the doors. For
other uses, the resistors can be made unequal in value to make some “votes” count
more than others, use a 2/3 majority, etc.

11.2.5 Example 5—Sine to Pulse Train Converter

Sometimes it is necessary to provide an on/off signal with a frequency determined by


a sine wave. For example, a digital alarm clock may get its timing from a sinusoidal
signal derived from the power outlet (i.e., 50 or 60 Hz, depending on location). Of
course, the sine wave input must be small enough in amplitude so that it does not
exceed the input ratings of the op-amp at any point in the cycle.
The circuit in Fig. 11.9 will produce a pulsed output between about 0 and 5 V
from a sine wave input. The “duty cycle,” which is the ratio of the “on” time to the
228 11 Non-linear Uses of Op-Amps

Fig. 11.9 Example 5—sine


wave to square pulse
converter

total time, can be varied by adjusting the relative values of R1 and R2 . If they are
equal, a 50% duty cycle would be expected. The feedback resistor, Rf , may not be
needed in all applications. It provides hysteresis to reduce false triggers, as described
above, and it is large compared to R1 and R2 . The lower diode shorts the output to
ground if the op-amp output goes negative and the diode to the +5 V supply shorts
the output to the 5 V power supply if the output of the op-amp exceeds 5 V. The
resistor at the output, R, should be large enough to keep the op-amp output current
within the range specified by the op-amp manufacturer.
For this application, an op-amp designed for 5 V, single-supply operation could be
used directly without the diodes and resistor R, though the resistor and diode protec-
tion may be necessary on the input to avoid excessive input currents, particularly
during the negative half of the cycle.

11.2.6 Example 6—Zero Crossing Detector

The circuit in Fig. 11.10 uses two comparators in order to provide a short pulse
when a signal crosses zero with a positive slope. It is easily modified to work with a
negative slope and/or to add an offset to detect when the signal crosses any desired
voltage level. The function of this circuit is the same as the trigger circuitry used
for an oscilloscope. The signals are illustrated in the figure are for a sinusoidal input
though any time dependent input can be used.

Fig. 11.10 Example 6—zero crossing detector, such as what might appear in an oscilloscope
triggering circuit
11.2 The Op-Amp Comparator 229

The first op-amp (U1 ) converts the input to a square wave—high if the input is
positive, low if it is negative. The capacitor and first resistor act as a high-pass filter,
and let through much more signal during the rapid changes in the square wave than
during the flat portions. The diode blocks the negative going pulses. The pulses at
this point have a sharp rise and a (rapid) exponential decay. The second op-amp (U2 )
compares the input to a reference voltage that is above zero, but possibly not above
zero by very much. The output of U2 then “squares up” the output, removing the
exponential decay.
If the diode is reversed, the circuit is sensitive to the negative slope. If an offset is
used, instead of ground, for the inverting input of U1 , then the trigger level will move
away from zero. Positive feedback can be added to provide hysteresis for either or
both comparators, as needed.

11.2.7 Example 7—Pulse Conditioner/Lengthener

Figure 11.11 illustrates a circuit useful for taking a short duration “messy” input
signal, and converting it to a nice “clean” square pulse, possibly significantly longer
in duration. The output signal might be used to signal that an event has occurred, such
as a click from a Geiger counter or the sound of a clap picked up by a microphone.
The circuit also has the feature that the strength of the signal can be reflected in the
length of the output pulse—a stronger input signal yields a longer output pulse. A
measurement of the pulse length is then a measure of the strength of the input pulse.
The first op-amp, U1 , is connected as an ideal diode circuit. Thus, when the input
is positive, the output should follow the input. The capacitor, C, will then charge (or
discharge) through R1 based on the amplitude of the input. The R1 C time constant
should be comparable to, or longer than, the expected duration of the longest input
pulse. Once the short-duration signal is complete, the capacitor can only discharge
through R2 . If R2 is much larger than R1 , the discharge time can be much longer than
the charging time. The second op-amp, U2 , is a simple comparator that outputs a
positive pulse as long as the voltage on the capacitor exceeds the reference level set
by R3 and R4 .

Fig. 11.11 Example 7—pulse conditioner/extender


230 11 Non-linear Uses of Op-Amps

11.3 Using the Comparator for Feedback

A comparator can be useful for feedback. A simple case would be a home heating
system where the furnace is to be either on or off depending on whether the house
temperature is above or below a preset value. A comparator of some sort checks the
temperature. If it is too high, the furnace is turned off and if too low, the furnace is
turned on. In such an application, some hysteresis would be used so that the furnace
is not turning on and off too often.
With some filtering, the comparator can be used for other feedback applications
leading to behavior that has the appearance of being continuous, rather than on/off.

11.3.1 Automatic Gain Control Amplifier

Figure 11.12 shows a purely electronic example of a comparator used for feedback.6
This circuit is designed to have a sinusoidal input and a sinusoidal output, but where
the output amplitude is (almost) independent of the input amplitude. Such a circuit
would be useful if there is a signal of unknown (or variable) amplitude and it is
desirable to have a “reference signal” at the same frequency, but with a fixed amplitude
(e.g., for a lock-in amplifier). This idea might also be used as an automatic amplitude
control for audio signals.7
Assume the op-amps are powered from ±12 V and V in is sinusoidal. The upper
op-amp (U1 ) functions as a linear amplifier with a nominal gain of ~100. The lower
op-amp (U2 ) compares the output from the first to a DC reference voltage, in this
case approximately 1.1 V. Any time the output from the first op-amp exceeds 1.1 V,
the output of the second op-amp goes from about −12 V to +12 V.
The diode D2 blocks the comparator output when the comparator output is −12 V.
In the absence of current through the diode, the capacitor connected to the gate of
the FET will charge to supply −12 V to the gate of the FET, which is more than
enough to pinch-off the FET. With the FET pinched-off, it is effectively removed
from the circuit and so the input is amplified by ~100. However, if the comparator
goes positive because some part of the signal has exceeded 1.1 V, D2 conducts and
the input voltage at the gate increases until the FET “turns on.” When it is on, the
FET is being used as a voltage variable resistor (since V ds is kept small) and the
input voltage applied to the op-amp is reduced—that is, if the output amplitude gets
too large, the input is reduced to compensate. Diode D1 is included to protect the
FET gate from positive voltages and the capacitor on the output of U1 blocks any
DC offset voltages from getting to U2 .

6 Specific values and components are given for reference and to serve as a starting point as there is
some interdependence for some values and no simple formula that can be applied.
7 Note that loudness is a perceived quantity that depends on many factors in addition to peak

amplitude, and so use as an automatic volume control may be problematic.


11.3 Using the Comparator for Feedback 231

Fig. 11.12 An automatic gain control amplifier for sinusoidal signals. Feedback derived from the
output of a comparator is used to keep the amplitude of the output signal constant for a large range
of input signal amplitudes

Fig. 11.13 The


experimental input/output
relationship measured for the
circuit of Fig. 11.12. Note
that the horizontal axis is
logarithmic

Measured values obtained with this circuit and a sinusoidal input are shown in
the graph in Fig. 11.13. The output amplitude is indeed quite constant as the input is
changed over a range of a bit more than a factor of ten.
232 11 Non-linear Uses of Op-Amps

11.4 Putting Pieces Together

Many tasks in electronics are accomplished using a combination of circuits, each of


which has a simpler, well-defined task to perform. The overall function requires they
each do their job and work together. Here is one example.

11.4.1 Simple Phase Sensitive Detector

The circuit in Fig. 11.14 shows (schematically) how one might construct an inexpen-
sive phase sensitive detector (a “lock-in amplifier”) by combining op-amp circuits
from this and the previous chapter.
If you trace through this circuit, the output (V out ) is a very low frequency signal
(nominally DC) that is proportional to that part of the input signal (V in ) which has
the same frequency as, and a certain phase relative to, a reference signal. Signals that
are at a different frequency or are out of phase will exit the multiplier as positive as
often as they do negative and will then average to zero in the low-pass filter.

Fig. 11.14 Op-amp circuits with different functions can be combined to create more complicated
instrumentation. Here, a design for a simple phase-sensitive lock-in amplifier is shown
11.5 Problems 233

11.5 Problems

1. Assuming the magnitude of the op-amp power connections are larger than V ref ,
predict V out as a function of V in for the comparator circuit in Fig. 11.P1. All the
resistors have the same value and ideal diodes may be assumed. Note that there
is no connection between V in and the resistors on the left even though the wires
cross in the diagram.

Fig. 11.P1 Problem 1

2. Describe the behavior of the output for the circuit in Fig. 11.P2 when the push-
button switch is pressed (closed) and then after it is released (opened) at some later
time. What happens if the button is pressed multiple times in rapid succession,
say once every 0.1 s for 1 s?

Fig. 11.P2 Problem 2

3. For the circuit in Fig. 11.P3, when do the LEDs turn on/off after the switch
has been pressed (closed) for a while and then released (opened) at t = 0? The
op-amps are powered with ±12 V.
234 11 Non-linear Uses of Op-Amps

Fig. 11.P3 Problem 3

4. A comparator with hysteresis is constructed following the schematic shown in


Fig. 11.3, using V ref = 0 V and Rr = 10 k. The measured output is shown in
Fig. 11.P4. What value of Rp was used?

Fig. 11.P4 Problem 4

5. A triangle wave with a peak-to-peak amplitude of 2 V is used for the non-inverting


input to a comparator. Sketch the output of that comparator for the three cases
where the inverting input is −0.5, 0, and +0.5 V. This illustrates a voltage to
pulse length conversion.
Chapter 12
Digital I

Digital electronics uses circuitry that has two possible states for inputs and outputs.
You might call them On and Off, 0 and 1, or maybe True and False. The mathematics
of logic, known as Boolean algebra, is based on the latter and is useful for digital
circuit analysis, at least as a starting point. In this chapter, this logic-based under-
standing of digital circuits is introduced followed by the introduction of flip-flops,
a first example in sequential logic where the output depends on both the inputs and
the previous state of the system.

12.1 Boolean Algebra

In summary, Boolean algebra1 includes the following:


• Elements that have one of two values—that is, they are “binary.” Here True = 1
and False = 0 will be used as the two values, but other choices are possible.
• A bar over the top of a value means “the other value” (a “NOT” when using True
and False). That is 1 = 0 and 0 = 1. The “NOT” operation acts on a single value.
• Two operations are defined for pairs of values:
OR, designated with a plus sign (+), returns false if both the values are false,
otherwise true.
AND designated with multiplication dot2 (·) which returns true if both the
values are true, otherwise false.

1 Boolean Algebra was named for the English Mathematician George Boole who first used it in the
mid-1800s as a system for the analysis of logic, long before digital electronic circuits existed.
2 As with multiplication, the dot is often omitted if the meaning is clear.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 235
B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_12
236 12 Digital I

Table 12.1 Truth table for


Input values Result
OR and AND
A B A+B A·B
0 0 0 0
0 1 1 0
1 0 1 0
1 1 1 1

Note that since the result of the NOT, AND, and OR operations belong to the same
set of values as the original arguments (i.e., true or false), the result of an operation
can be used as an input for another operation, a subtle but important point.3 Note
also that the OR and AND operations are commutative, associative, and distributive.
By convention, when an expression contains multiple operations, the order of
execution for the operations is similar to what is used for addition and multiplication.
Quantities in parentheses are evaluated first, then ANDs followed by ORs. When
a NOT bar covers an expression involving several values, the expression is to be
evaluated first, before the “NOT” is applied. As examples, consider the following:
• A + B · C → The AND is applied to B and C, and that result is ORed with A.
• A · (B + C) → The OR is applied to B and C, and that result is ANDed with A.
• A · B + C → The AND is applied to A and B, then the NOT is applied to that
result. The result of that NOT is ORed with the result of the NOT applied to C.
• A + B · C → The OR is applied to A and B, then the (lower) NOT is applied to
that result. The result of that NOT is ANDed with C, and then the (upper) NOT
is applied to that result.
Notations that result in possible ambiguity should be avoided. For example, A ·
B + C might be confusing.
( Adding
) parentheses often helps. For example, the order
of operations for A · B + C is unambiguous.
Since there are only four possible input combinations each for the AND and OR
operations, it is not hard to write them out in a “truth table,” such as Table 12.1.
Truth tables are an alternate and simple way to show all the possible inputs and the
corresponding outputs.
The double operations obtained by applying a NOT to the result of both AND and
OR are seen in electronics often enough they are given their own names—NAND
and NOR respectively.4 For the NAND operation the output is false only if all the
inputs are true, and for NOR the output is true only if all inputs are false.

3An example from physics where this does not occur is the vector dot product.
4Even though the NOT is applied last, the “N” representing the NOT operation is put at the beginning
of the name. Possibly this is because ANDN and ORN are much less pleasant to pronounce.
12.3 Digital Logic Circuits 237

12.2 Useful Rules and Theorems for Boolean Algebra

A number of useful relations for Boolean algebra are listed in Table 12.2. Many of
these are relatively obvious. The last two entries are known as DeMorgan’s theorems
and are very important for the understanding and design of digital circuits. It is
important to become proficient with the result of DeMorgan’s theorems. All of the
relations in Table 12.2 are easily proved using truth tables. A truth table shows all
of the possibilities and if they all agree then the theorem is rigorously proved. For
example, the proof for one of DeMorgan’s theorems is shown in Table 12.3—the
middle and last columns are the same for all possible input combinations, so they
are equivalent. Such a proof may not seem elegant, but it is rigorous.
There are sixteen possible unique operations that can be defined that combine two
binary values (A and B) to give a binary result. Some of these operations are trivial,
but all of them can be expressed using the NOT, OR, and AND operations above.
Hence, the Boolean operations form a complete set.5 Table 12.4 shows all sixteen
operations. Note that for the first six entries, the value of one or both of the arguments
is immaterial—the result of the operation does not care.
The last two entries in Table 12.4 involve the so-called “exclusive or” (XOR). The
exclusive OR more closely matches the use of the word “or” in everyday language.
The result is true if either A or B is true, but not both. The usual Boolean OR,
sometimes referred to as the “inclusive OR,” will be true if either is true, including
the case where both are true. For convenience, the special symbol “⊕” is sometimes
used for the XOR function, though that operation is not really a part of the algebra.
Be careful using XOR with Boolean algebra because XOR does not have all of the
nice properties you might expect for a mathematical operation.
Where possible, the logic operations are usually pronounced as single words.
XOR is pronounced with two syllables (“ex-or”). To be consistent, XNOR might be
written NXOR, but that is too hard to pronounce—hence “ex-nor,” with the X first,
is used for the sake of pronounceability.

12.3 Digital Logic Circuits

Digital logic is implemented electronically using circuits designed to perform these


Boolean operations. Hence, those electronic components are often referred to as
“logic circuits” or “logic gates.” Schematic symbols for the circuits that perform
the basic operations are shown in the Table 12.5. As shown in this table, the inputs
are supplied on the left and the output is on the right. Many of these circuits are
also available with more than two inputs, in which case the description on the right

5 In fact, one can write all of these operations using only the NAND or only the NOR operations.
Hence, there is some redundancy to the operations. In terms of digital circuit design, this means
there will often be many different ways to obtain the same result.
238 12 Digital I

Table 12.2 Some basic relationships found in Boolean algebra

A=A A+B = B+ A
A+0=0+ A = A A·B = B· A
A·0=0· A =0 A+ A·B = A
A+1=1+ A =1 A + (B + C) = ( A + B) + C
A·1=1· A = A A · (B · C) = ( A · B) · C
A+ A= A A · (B + C) = A · B + A · C
A· A= A A + (B · C) = ( A + B) · (A + C)
A+ A=1 ( A + B) = A · B
A· A=0 A·B = A+B

Table 12.3 Truth table proof for one of DeMorgan’s theorems—for all A and B, A + B = A · B
A B A+B A+B A B A·B
0 0 0 1 1 1 1
0 1 1 0 1 0 0
1 0 1 0 0 1 0
1 1 1 0 0 0 0

Table 12.4 The set of


Name Description of result
Boolean operations that are
possible for two inputs Null or false 0 (no matter what)
One or true 1 (no matter what)
A A (i.e., B is ignored)
B B (i.e., A is ignored)
NOT A A (B is ignored)
NOT B B (A is ignored)
A AND B A·B
A OR B A+B
A AND (NOT B) A·B
(NOT A) AND B A·B
NAND of A and B A·B
A OR (NOT B) A+B
(NOT A) OR B A+B
NOR of A and B A+B
XOR of A and B A·B+B· A
XNOR of A and B A·B+ A·B
12.3 Digital Logic Circuits 239

Table 12.5 Basic digital gate


Name Schematic Logic function
schematics
Buffer Output equals input

Inverter Output opposite of input (NOT)

OR Output F only if all inputs F

NOR Output T only if all inputs F

AND Output T only if all inputs T

NAND Output F only if all inputs T

XOR Output T only if just one input T


(Is F if both inputs same)
XNOR Output T only if both inputs same

applies. That is, an “eight-input NAND” circuit would produce F (false) only if all
eight inputs were T (true).
Notice the convention that in the schematic the “NOT” function is added using
a small circle. A small circle on the input side would indicate that the “NOT” is to
be performed on the input before the logical operation is performed. For example,
consider the three schematics in Fig. 12.1 that have corresponding outputs A ⊕ B,
A ⊕ B, and A ⊕ B.
For digital logic circuits, the two logic states are associated with two voltages
and/or currents. Most common is the use of 0 and 5 V or 0 and 3.3 V, with 0 V being
“false” in both cases. More accurately, any voltage smaller than a certain maximum
is considered “low” and any value higher than a certain minimum is considered
“high.” Thus, some truth tables will be written using H and L, rather than 1 and 0,
corresponding to those voltage levels. Between high and low is a region where results
are unpredictable. Digital logic circuits are produced in “logic families” that have
matching electrical characteristics. For details on any particular circuit, refer to the

Fig. 12.1 Three examples of how the NOT can be shown in a schematic using a small circle. The
NOT can be applied to the input or output. The three schematics correspond to the operations A ⊕ B,
A ⊕ B, and A ⊕ B, respectively
240 12 Digital I

Fig. 12.2 Three examples of how simple digital circuits can be constructed using discrete diodes
and transistors, a the NOT, b the NAND, and c the NOR

manufacturer’s datasheet for the device. When combining digital logic circuits, it is
safest, though not always necessary, to stick to circuits from the same logic family.
Examples of one way to create logic functions using diodes and transistors is
shown in Fig. 12.2. In Fig. 12.2a, a high input turns on the transistor, and so the
output is low and a low input leaves the transistor off, so the output is high. That
is the NOT function. In Fig. 12.2b if either or both of the inputs are low then the
transistor is off, and the output will be high. This is the NAND function. And in
Fig. 12.2c, if either input is high, the transistor is on, so the output is low. This is the
NOR function. This diode-transistor logic (DTL) is easy to construct with discrete
components, however its use within integrated circuits is very rare.
Note that power supply connections are rarely shown for digital logic circuits,
but are necessary to make the circuit function. Logic that uses voltages levels will
typically use those same levels for the power connections. That is, if logic levels are
0 and 5 V, then the power supply connections are also 0 and 5 V. When power supply
connections are shown on a schematic, they are often shown separate from the logic
circuitry. The same power supply connections may be shared by several logic gates.

12.4 Combinations of Digital Logic Gates

When logic gates are connected together, the resulting output can be determined using
Boolean algebra or by writing out a truth table. Three examples follow demonstrating
how this is accomplished.

12.4.1 Example 1—Solving with Boolean Algebra

Consider the circuit of Fig. 12.3. Boolean algebra can be used to figure out what it
does. When solving, it often helps to include some of the intermediate states, as is
shown in the figure. Also observe that this particular circuit is symmetric so that it
12.4 Combinations of Digital Logic Gates 241

Fig. 12.3 Circuit for


Example 1. Intermediate
values have also been labeled

should be the case that swapping the inputs A and B should not change the output
value S.
Consider the inputs to the last NOR gate and (carefully) make use of DeMorgan’s
theorems:

C = A + A + B = A · (A + B) = A · (A + B) = A · A + A · B = A · B.
(12.1)

Similarly,

D = B · A, (12.2)

and so, again using DeMorgan’s Theorems:

S =C+D = A·B+B· A= A·B·B· A


( ) ( )
= A + B · B + A = A · B + A · B. (12.3)

The final result is that of the XNOR operation—the output is true (i.e., 1) if both
inputs are the same, and is false (i.e., 0) otherwise. Note that the symmetry predicted
is indeed present in the answer.
Suppose there were three inputs. What circuit could be used to check to see if
all three inputs are the same? How might such a circuit be constructed using only
2-input NOR logic gates? (Consider the statement “if A and B are the same, and B
and C are the same, then A, B and C are all the same.”).

12.4.2 Example 2—Solving with a Truth Table

Consider the digital circuit in Fig. 12.4. To figure out what it does, create a truth table.
The intermediate value “D” is defined and is included in the table for convenience.
Recall that each exclusive OR gives a true when its inputs are different, and a false
when they are the same. Table 12.6 shows resulting truth table, worked out one row
(i.e., one set of inputs) at a time.
242 12 Digital I

Fig. 12.4 Circuit for


Example 2

Table 12.6 Truth table for Example 2


Inputs Result
A B C D=B⊕C S=A⊕D
0 0 0 0 0
0 0 1 1 1
0 1 0 1 1
0 1 1 0 0
1 0 0 0 1
1 0 1 1 0
1 1 0 1 0
1 1 1 0 1

It is always desirable, though not always possible, to try to put what the circuit
does into simple words. Such a description may not be unique. One description for
this circuit would be that “the output is 1 if the number of non-zero inputs is odd,
and 0 if even.” Whether or not the number of 1’s in a sequence of binary digits is
even or odd is known as the “parity.” Hence, this circuit tells you if the parity is even
(S = 0) or odd (S = 1) for these three inputs. How can this circuit be generalized to
determine the parity of N inputs?

12.4.3 Example 3—Solving Both Ways

It may not be obvious what function is performed by the digital circuit in Fig. 12.5,
however it is straightforward to create a truth table. Since there are three inputs,
there should be 23 = 8 rows (plus the header). Intermediate values (D, E, and F) are
included to help reduce errors. The resulting truth table is shown in Table 12.7. It can
be seen that G, the result, is the opposite of C for all rows except the second, which
is one of the two rows where both A and B are 0. Or to phrase it another way, the
output is true if A and B are both false or if C is false. Thus, translating those words
into the algebra, the solution should be.

G = C + A · B = C + A + B. (12.4)
12.4 Combinations of Digital Logic Gates 243

Fig. 12.5 Circuit for


Example 3

Table 12.7 Truth table for Example 3


Inputs Result
A B C D E F G
0 0 0 1 0 1 1
0 0 1 1 1 0 1
0 1 0 0 1 1 1
0 1 1 0 0 1 0
1 0 0 0 0 1 1
1 0 1 0 1 1 0
1 1 0 0 1 1 1
1 1 1 0 0 1 0

The result can, of course, also be derived using Boolean algebra. In fact, one good
way to ensure a solution is correct is to do it both ways. The result better be the same.
The algebraic steps for this example are

D = A + B = A · B; E =C · B +C · B (12.5a)

( )
F = D· E = A· B · C · B +C · B
( ) ( )
= A· B · B ·C + A· B · B ·C = A· B ·C (12.5b)

G =C · F =C · A· B ·C

= C · ((A · B) + C) = C · (A + B) + C · C
=C+ A+B =C+ A·B (12.5c)

DeMorgan’s theorems were used numerous times, as well as some other identities
from Table 12.2. A particular challenge for this example is keeping careful track of the
operation order and the NOTs inside NOTs inside NOTs. The final result suggests that
the same function could have been obtained using a much simpler circuit. Boolean
244 12 Digital I

algebra can be a useful tool to search for alternate and/or simpler designs that perform
the same logical function.6

12.5 Equivalent Circuits

With digital circuits, there is always more than one way to achieve a desired function.
This can be seen by considering the examples in Fig. 12.6 constructed entirely using
2-input NAND gates, but that produce other (simple) functions (NOT, OR, AND).
Since all of the basic operations can be constructed entirely from 2-input NANDs,
any digital circuit can, in principle, be constructed solely from 2-input NANDs.
As another example consider the construction of the XOR function. The basic
relationship for XOR is

A ⊕ B = A · B + A · B. (12.6)

The circuit of Fig. 12.7a follows immediately from the right-hand side. Other
equivalent circuits can be discovered using Boolean algebra. One “trick” to use is
that any expression that is always false can be ORed with another expression without
changing its value, in the same way zero can always be added to a numerical equation.
Starting with that idea, then using the distributive property and one of DeMorgan’s
theorems, an alternate XOR circuit is found as follows:

A⊕B= A· B+ B· A = A· A+ A· B+ B· A+ B· B
( ) ( )
= A· A+B +B· A+B
( ) ( )
= (A + B) · A + B = (A + B) · A · B , (12.7)

that is constructed as shown in Fig. 12.7b. It is also possible to make an XOR in


many other ways. Suppose it is desired to make an XOR using only 2-input NORs.
Start the algebra as before but work to turn all ANDs into ORs using DeMorgan’s
theorems. Then make sure there is at least one bar over them for the NOT. That is,
starting the algebra as before,
( )
A ⊕ B = (A + B) · A + B
( )
= (A + B) · A + B = A + B + A + B. (12.8)

The far right-hand side can be implemented using only NORs. In fact, all the basic
operations can be achieved using 2-input NORs, and if desired, any digital circuit
can be constructed solely with NORs.

6 Another method to simplify logic circuits uses a so-called Karnaugh map. It is easy to find many
articles about how to use these maps.
12.6 Gates Versus Logic Functions 245

Fig. 12.6 Four examples showing how NAND circuits can be used to produce other basic functions.
From top to bottom the results are NAND, NOT, AND, and OR

Fig. 12.7 Two examples showing the construction of the XOR function using basic Boolean
operations

12.6 Gates Versus Logic Functions

With digital circuits, a single circuit can appear to have radically different functions
when viewed in context. Consider the examples in Fig. 12.8. The first (Fig. 12.8a)
is a simple logical combination of two inputs that appear to be treated equally. The
same circuit is shown in Fig. 12.8b, but labeled with a “control input” (cntrl) that
either allows D to get through (when cntrl is true, then D' is equal to D) or forces the
output to be 0 (when cntrl is false), thus blocking the data D from getting through. In
this context, the inputs may not appear to be treated equally. A simple XOR circuit
is shown in Fig. 12.7c where D ' = D if cntrl is 0, and D ' = D if cntrl is 1. When
used in this way, the XOR is called a “controlled NOT,” or CNOT gate. The quantum
CNOT gate is very important for quantum computing.
246 12 Digital I

Fig. 12.8 The interpretation of the function of a circuit can depend on context. At a is a simple
logic application. In b the same circuit appears to be used for control functions. In c the XOR is
being used as a controlled NOT

When presented as in Fig. 12.8b, c, the digital circuits are acting like gates that
either let the data through, block it, or invert it. When used in such a manner, the
“logic” does not seem as important as does the control function, though the electronics
is identical. In practice, logic gates are most often, in fact, used to perform functions
that are not based in logic, although they can be analyzed using logic. After all,
digital logic gates are usually wired to perform some specific (electronic) function,
not to ponder philosophy.

12.7 Decoders and Encoders

Decoders and Encoders are often used to efficiently send data from one place to
another, usually for control, input or display purposes. Decoders take a combination
of inputs and produce a desired outcome based on that combination.
An example of an “active line low” decoder is shown in Fig. 12.9 and the function
is shown in the truth table, Table 12.8. The indices have been chosen so that the two
inputs can be combined mathematically to get a label (i.e., 2 × B + A) corresponding
to the output that is forced low. Thus, the signals on two wires can be decoded to
select one of four output lines.
While the circuit and wiring may get somewhat more complicated it is not difficult
to increase the number of inputs. Consider such a decoder with n inputs. Since there

Fig. 12.9 Circuit for a


simple decoder
12.7 Decoders and Encoders 247

Table 12.8 Truth table for decoder circuit


B A O0 O1 O2 O3 2B + A
0 0 0 1 1 1 0
0 1 1 0 1 1 1
1 0 1 1 0 1 2
1 1 1 1 1 0 3

Fig. 12.10 Numerical digits are often displayed using a 7-segment scheme. A decoder circuit will
take a combination of four inputs and will activate the appropriate segments for each digit

are 2n possible combinations of inputs, 2n outputs can be used. The advantage for
large n is obvious—with 10 wires you can control the output on 210 = 1024 wires.
Another type of decoder is used for displaying numbers using a 7-segment display.
Refer to Fig. 12.10. Here there are four inputs7 and 7 outputs. Six of the sixteen
possible input combinations are not needed for this task and may or may not produce
a meaningful display. The seven output lines are used to selectively turn on one of
seven bars of the display. Should it be necessary, the first step to build such a decoder
is to look at each segment in the display, one at a time, and write out a truth table for
that segment. In some cases, it might be better to determine when the segment is off,
rather than on. For example, segment a is on for all digits except 1 and 4, segment
b is on for all except 5 and 6, and so on. Thus, it might be simpler to construct the
logic to turn segments off, rather than on. The internal workings for such a decoder
will not be reproduced here. Fortunately, inexpensive integrated circuit 7-segment
decoders can be purchased that perform this function.
An encoder performs the inverse function from that of a decoder. The truth table
will look similar to that for the decoder, but with inputs and outputs swapped. A
simple 4-input encoder is shown in Fig. 12.11 with its truth table in Table 12.9. In
this case a high (1) on one of the inputs causes a corresponding pair of binary values
to appear on the outputs. An additional output, S, is included to indicate that one (or
more) of the input lines is high—S provides an indication that an input has occurred.
To shorten the truth table, and to more easily understand the circuit, the symbol
“X” is introduced to indicate a value that can be either 0 or 1 with no change in the
output (given the other input values). That is, for the circumstances shown in that line
of the truth table, that particular input does not matter. The “X” is sometimes referred
to as “Don’t Care.” With that shorthand, the sixteen possible input combinations for
the encoder can be written using many fewer rows, and the table is easier to read and
understand.

7 Since there are 10 digits, and 3 inputs yields only 8 possibilities, there must be at least 4 inputs.
248 12 Digital I

Fig. 12.11 A circuit for a


four-input encoder, where
the two output lines will
contain a code for which
input line is high

Table 12.9 Truth table for 4-line encoder


a3 a2 a1 a0 b1 b0 S 2b1 + b0
1 X X X 1 1 1 3
0 1 0 X 1 0 1 2
0 0 1 X 0 1 1 1
0 0 0 1 0 0 1 0
0 0 0 0 0 0 0 0
0 1 1 X 1 1 1 3

One possible problem with this simple encoder occurs when two or more of the
inputs are high at the same time. The corresponding decoder had only four possible
states, but this encoder has sixteen. For this example, in all but one case the output
with the larger index will appear at the outputs. That is, if a3 and a1 are high together,
a1 is ignored and the result is as if only a3 were high. With a small amount of additional
circuitry, the one remaining case (if a1 and a2 are simultaneously high, as seen on the
last line of the truth table) can be “fixed” so that the output always represents that for
the highest value index. Such an encoder that, by design, resolves all such conflicts
in one way or another would be referred to as a “priority encoder.”
One use of an encoder/decoder combination would be the efficient transmission
of data. For example, 2N separate inputs can be compressed into N data lines for
transmission, then turned back into 2N separate outputs at the other end. Decoders
and/or encoders can also be used in pairs to create a “row/column” set-up used in
“multiplexing.” As will be seen in the next example, this is an efficient way to wire
(for example) some displays. An example showing how this is used for a modern
keyboard is shown later.

12.8 Multiplexing

Multiplexing allows many devices to use some common electronics and can result
in simpler circuitry and significantly lower cost.
12.9 Flip-Flop Circuits 249

Fig. 12.12 Two decoders in


a row-column arrangement
can be used to choose one of
sixteen LEDs to illuminate
based on four input lines

Consider the circuit in Fig. 12.12 used to illuminate 1 of 16 LEDs. In this


schematic, the convention used is that wires that cross are not connected unless
there is a solid circle (a ‘blob of solder’) at the crossing point.
For this circuit, four values determine the row and column of the LED to illuminate.
The two decoders are functionally the same as the one described in Table 12.8, though
one is wired for selected line high rather than low. For any combination of the four
inputs, only one LED will have a high level on the anode (the row) and a low level
on the cathode (the column), and that one LED will light up. The resistors (which
may not be necessary in some circumstances) limit the current (typically 5 to 15 mA
for normal LEDs). While only one LED can be illuminated at a time, if multiple
LEDs are illuminated in rapid sequence, they will appear to a human observer to be
simultaneously illuminated.
The row and column design make it possible to “address” 2N things (in this case
LEDs) using N inputs. While the savings going from 4 to 16 may not seem so
impressive, imagine if N = 16, requiring 16 inputs (each with a wire) that allows
216 = 65,536 things to be addressed that might be, for example, “64 k” memory
locations—a trivial amount of memory but far from a trivial number of wires.8

12.9 Flip-Flop Circuits

The circuit of Fig. 12.13a seems to be showing a form of circular logic. There are
two inputs, R and S, and two outputs Q and Q' , however the outputs are also being
used as inputs. This circuit is usually presented as shown in Fig. 12.13b, where the
lower NOR gate has been twisted around so all the inputs are on the left and the
outputs on the right.

8In this context, the number 1 k is 210 = 1024, that is close to, but slightly larger than, 1000. Hence,
64 k is also slightly larger than 64,000.
250 12 Digital I

Fig. 12.13 A simple RS flip-flop is shown in (a) where circular logic is clearly shown. The circuit
in b is identical to that in (a), except all the inputs have been moved to the left. In c an extra control
input, the gate, is added to make a gated RS flip-flop

Table 12.10 Truth table for


R S Q Q'
Fig. 12.12a
1 0 0 1
0 1 1 0
1 1 0 0
0 0 ? ?

A NOR will produce a false (0) if any input is high. Hence, if R is high, then Q
is low, and if S is low, then Q' is high. Likewise, if S is high, Q' is low, and if R is
low, Q is high. The remaining state, when both R and S are low, is where it becomes
interesting. Consider what happens for that case if Q is high. In that case Q' must
be low, which makes Q high, as was assumed. All is well. However, if instead it is
assumed that Q is low, then Q' must be high, which forces Q to be low. All is well
with the opposite assumption also. So, is Q high or low? Logically, there are two
valid, though contradictory, solutions. The resulting unresolved truth table is shown
in Table 12.10.
A solution for a real electronic circuit is found by noting that before R and S were
both low, they were something else. If those previous values were 0 and 1, or 1 and 0,
the solution remains valid when they both R and S become 0. Hence, the outputs stay
what they were before. The previous state is stored. The case where both R and S are
1 followed by both R and S 0 leads to unpredictable results. In any real circuit, both
will not switch simultaneously—even a few nanoseconds of difference is enough—
and even then the output state with both Q and Q' zero is not a stable solution. The
circuit will end up in one of the stable states. It is not obvious which one. Hence,
going from both inputs true directly to both inputs false should be avoided.
Circuits such as in Fig. 12.13a, b, that have two stable states that can be stored,
are known as flip-flops. This particular circuit is known as an RS flip-flop. If you
think of Q as being the principal output, then S serves to set the value (to 1) and R
serves to reset the value (to 0). Aside from the unstable situation when both R and S
are 1, Q' is always the opposite of Q. While RS flip-flops are not all that common as
separate devices in circuits, they serve as the building block for flip-flop circuits that
are. Flip-flops are an example of sequential logic, in that the output does not only
12.9 Flip-Flop Circuits 251

depend on what the inputs are now, but also on the sequence of events that occurred
before “now.”
With the addition of a control function out front, the RS flip-flop can be enabled
and disabled (with a “gate” or “clock” signal). In that case, the set and reset functions
will only occur if the gate signal allows them through. In practice, the gate will be
used to synchronize actions to an external signal. The inputs (R and S) are set to their
desired values before the gate is opened, then the gate opens and lets them through
at a prescribed time. On the other hand, an overall set or reset might be desired that
should override such a gate, so extra inputs can be brought out for those purposes.
Figure 12.13c shows a “gated RS flip-flop with preset and clear” constructed with
NANDs. The preset and clear functions set the output to 1 or 0. Note that in this case
the functions (or their abbreviations) have a bar over them in the diagram. The bar
indicates that a low (0 or false) input preforms the desired action. For example, to
preset the flip-flop, a low level is applied to the P R input—if P R is made false, then
PR is true. It is a funny way to use negatives, to be sure, but that is a standard practice.
The corresponding truth table for this circuit is shown in Table 12.11. The symbol
Q0 here is used to designate the previous value of Q—that is, the starting value of
Q when the current state began. There are still some states that are “unstable” and
should be avoided.
Two possible solutions to deal with the unstable states of the RS flip-flop are
the so-called JK flip-flop and the D flip-flop. The JK flip-flop ultimately creates an
alternative output to replace the unstable state, while the D flip-flop limits the inputs
so the unstable state cannot be reached.
A JK flip-flop is shown in Fig. 12.14a. The circuit may also have preset and clear
functions that are not shown. The JK flip-flop is essentially the gated RS flip-flop
with the addition of the connections between the outputs and the inputs. Here J is
the set and K is the reset input. The feedback from the outputs is such that the set
function is disabled if the circuit is already set, and the clear function is disabled if
it is already cleared. Hence, only one input will be effective at a time. An interesting
thing happens if both inputs are true—since the only input that is active is the one
that does not match the current output, the output will switch, or “toggle” to the other

Table 12.11 Truth table for Fig. 12.12c


PR CL Gate S R Q Q'
0 1 X X X 1 0
1 0 X X X 0 1
1 1 0 X X Q0 Q0
1 1 1 1 0 1 0
1 1 1 0 1 0 1
1 1 1 0 0 Q0 Q0
0 0 X X X Unstable
1 1 1 1 1
252 12 Digital I

Fig. 12.14 Extensions of the RS flip-flop include the gated a JK flip-flop and b the D flip-flops.
An alternative D flip-flop that uses fewer logic circuits is shown in (c)

state. The truth table for this JK flip-flop is shown in Table 12.12. There is a problem
in that the gate can only be kept high for a very short time, otherwise the output
will continue to toggle back and forth as long as the gate is open. This issue will be
resolved shortly.
A D flip-flop is shown in Fig. 12.14b. Here the two inputs, R and S, are replaced
with one input, and opposite values are applied to the two inputs of the RS flip-flop,
forcing R = S. The unstable state cannot happen and it will always be the case that
Q ' = Q. The D flip-flop is a basic storage element for data (D). The truth table for
the D flip-flop is shown in Table 12.13. An alternative, and somewhat simpler circuit
that performs the same function is shown in Fig. 12.14c. Preset and clear functions
can be added to either of these circuits, if desired.

Table 12.12 Truth table for Fig. 12.13a


Gate J K Q Q'
0 X X Q0 Q0
1 0 0 Q0 Q0
1 1 0 1 0
1 0 1 0 1
1 1 1 Q0 Q0

Table 12.13 Truth table for


Gate D Q
Fig. 12.13b, c
0 X Q0
1 0 0
1 1 1
12.10 Edge-Triggered Flip-Flops 253

Fig. 12.15 Two gated D flip-flops in series, with an inverted gate input, can produce an edge-
triggered D flip-flop, where the output is loaded only when the gate changes

12.10 Edge-Triggered Flip-Flops

A very useful device is constructed using two flip-flops in series. This will be illus-
trated using the D flip-flop, though a similar result can be obtained using other flip-
flips. The details of the construction of the flip-flop are not important at this point,
so the D flip-flop is represented by a simple box with the inputs and outputs labeled
appropriately. The two flip-flops in series are connected as shown in Fig. 12.15. Due
to the inverter (NOT), only one of the flip-flops can have its gate open at any given
time.
Each of the D flip-flops have the property that while the gate is open, the output
(Q) follows the input (D). If the gate is closed, the output is the stored value of D
that was the last to occur before the gate closed. For the circuit of Fig. 12.15, when
the clock is high, the first flip-flop has its gate open and its output follows the input.
That output, Qi , is presented to the second flip-flop, and that flip-flop ignores the
value since its gate is closed. When the clock goes low, the first flip-flop stores and
continues to output the last value it saw at the input. At the same time, the second
flip-flop now opens its gate, and so its output follows its input. Since the second
flip-flop’s input is now the stored value from the first flip-flop, the output from the
second flip-flop becomes the value of the input at the time the clock changed from
high to low. If the clock goes back high, the output from the second flip-flop does
not change. Only a transition from high to low can affect a change.
The truth table, that is, a description of what happens, for the circuit of Fig. 12.15
is shown in Table 12.14. Here the down arrow, ↓, indicates a change in value from
high to low. Such a device is referred to as “edge-triggered.” In contrast, the original
flip-flop might be referred to as “level-triggered.” It is certainly possible to modify the
circuit to trigger on the rising edge (↑).9 An edge triggered flip-flop appears to grab
the input at a particular time, though internally it is really a hand-off between two flip-
flops. An edge triggered JK flip-flop is constructed in a similar manner, eliminating
the problem of having long gate signals when in toggle mode, mentioned above.
When the internal construction is not of concern, an edge-triggered D Flip-flop
would appear in a schematic something like what is shown in Fig. 12.16. That it
has an edge-triggered input is indicated with a small triangle. The circuit on the left

9 In a truth table, an image representing a falling or rising edge may be used instead of an arrow.
254 12 Digital I

Table 12.14 Truth table for


D Clock Q
Fig. 12.14
X 1 Q0
X 0 Q0
1 ↓ 1
0 ↓ 0

Fig. 12.16 Schematic representations of edge-triggered D flip-flops. At a, the rising edge is used
and in b the falling edge. Such flip-flops may have additional inputs for preset and clear functions

(Fig. 12.16a) is triggered on the rising edge (low to high or “going true”) while the
circuit on the right (Fig. 12.16b) is triggered on the falling edge (high to low or “going
false”), as indicated by the “not” circle on the input. If the device also has preset,
clear, or other functions, they are simply added to the box with appropriate labels.
These edge-triggered flip-flops are available in various configurations as integrated
circuit devices.

12.11 A Directional Electric Eye

A simple electric eye detects the presence of an object between a light source and
a light detector. The circuit of Fig. 12.17a uses two such detectors and an edge-
triggered flip-flop to also determine the direction of motion of the object. In this
circuit, the light detector is a phototransistor that conducts when there is incident
light. Hence, in the absence of an object the transistors are on and both comparators
will output a low. Assuming the object is larger than the spacing between the two
sensors, the output of the two comparators as the object passes will be as shown in
Figs. 12.17b or 12.17c, depending on direction. In Fig. 12.17b the object is traveling
up the page, while in Fig. 12.17c it is going down the page. The flip-flop will store
the value of the D input when the clock goes from low to high. In Fig. 12.17b, D will
be high, and in Fig. 12.17c, D will be low at the time when the clock goes high.
Such circuitry is common in the design of early mechanical computer mice and
trackballs. A wheel with periodic openings is placed between the sensors. The opaque
region between the openings acts as the object. Similar designs are still used with
some optical rotary (or shaft) encoders. Measuring the frequency of the pulses from
12.12 Combinations of Flip-Flops 255

Fig. 12.17 Two optical detectors and an edge-triggered flip-flip can be used to detect the passage
of an object and be able to determine the direction of motion

either of the two detectors, plus the output of the flip-flop, provides a measure of
the speed and direction of rotation. A much simpler application for this circuit is to
measure the period of a pendulum. As the pendulum swings through the detectors,
one pulse occurs (from the flip-flop) for each full swing.

12.12 Combinations of Flip-Flops

The D Flip-Flop is a very simple memory since it can store a single previous value,
a “bit,” indefinitely. An array of flip-flops can obviously be used to store more data.
In addition, there are a number of interesting circuits based on interconnected edge-
triggered flip-flops.

12.12.1 Shift Register

A particularly simple application, illustrated in Fig. 12.18, is the so-called shift


register. The output of each flip-flop is connected as the input to the next. When the
clock makes its transition, in this case low to high, the data moves one flip-flop to
the right. That is, Q4 changes to the previous value of Q3 , Q3 changes to the previous
value of Q2 , and so on. In addition to the applications discussed below, a shift has
important applications in serial communications and multiplication, both discussed
later. Shift registers are available as integrated circuit devices.
A shift register with feedback from the last stage forms a ring counter. If the
feedback is inverted, as shown in Fig. 12.19, then it is referred to as a Johnson
counter. If such a counter starts with all zeros (or all ones) and has n flip flops, it
will divide the clock frequency by 2n. Such a circuit has a potential for 2n different
output states, and is using only 2n of them. Provisions are usually included to ensure
that the counter starts in a known state (e.g., all zeroes), or if it gets into a state it is
not supposed to, it will be set to a state that is ok.
256 12 Digital I

Fig. 12.18 Edge-triggered D flip-flops connected in series, where the output of each is connected
to the input of the next, can form a shift register

Fig. 12.19 A shift register where the output of the last stage is connected back to the first forms a
ring counter. If the inverted output of the last stage is used, as shown, it is sometimes referred to as
a Johnson ring counter

A shift register with “taps” to provide its own input from a combination of several
of the later stages can generate useful sequences. These are examples of one type of
“linear feedback shift register” (LFSR). For example, using the correct taps XORed
(or XNORed) together from N flip-flops that are fed back to the beginning can lead
to a “maximal length sequence” of 2N − 1 distinct values having statistics similar to
those of random numbers. The sequence is completely determined and will repeat,
however. Such a sequence is referred to as “pseudo-random.” Correct tap positions
for such a sequence can be found in published tables.10 Figure 12.20 shows such
a circuit with N = 4 that will produce 15 pseudo-random values before repeating.
If a similar circuit with 31 flip-flops is constructed, with taps at the 24th and 31st
flip-flops, a sequence of just over 2 × 109 pseudo-random 0’s and 1’s is produced
before repeating.
The initial state of each flip-flop is set using a “seed” value using “preset” or “clear”
inputs on the flip-flops (not shown in the schematic). For the circuit of Fig. 12.19
the only bad seed is all ones, that simply reproduces itself indefinitely. Pseudo-
random number generators and maximal length sequences are an important compo-
nent of spread-spectrum communications used by GPS (and other) satellites, some
cell phones and other wireless technologies, as well as for many games. Random
values can also be used as a tool to simulate various random processes in the physics
lab.

10There are always an even number of taps, and, numbered starting from 0, the tap positions are
“setwise prime,” meaning no common factor, other than 1, divides all of them.
12.12 Combinations of Flip-Flops 257

Fig. 12.20 A shift register where the outputs of several stages are combined with an XOR or XNOR
can produce a pseudo-random number generator if the correct tap points are used. Such a circuit
with N flip-flops will cycle through 2N − 1 states before repeating

12.12.2 Binary Counter


( )
If an edge-triggered D flip-flop is connected so that the inverted output Q is wired
to its own data input, then with each clock pulse the output will toggle to the other
state.11 In Fig. 12.21 several such D flip-flops are connected so that the output (Q)
of each flip-flop is connected to the clock input of the next. Thus, when the output
Q0 goes from high to low, the output Q1 will toggle, and when Q1 goes from high
to low, Q2 will toggle, and so forth. Figure 12.22 illustrates the sequence of outputs
starting from all low (0’s). In the process, all 24 = 16 output states are encountered
for the four flip-flops, before the cycle repeats.
As a frequency divider, a circuit with N flip-flops can easily divide the clock
frequency by a factor of 2N . If clear inputs are also available, additional logic can
reset the values back to all zeros at any point, allowing division by any integer value
up to, and including 2N .
If a high is taken to be 1 and a low 0, then the number of pulses, n, that occurred
(since the outputs were last all zero) can be computed from the outputs using

n = Q0 + 2 × Q1 + 4 × Q2 + 8 × Q3
N −1
E
= 2k Q k . (12.9)
k=0

For example, at the time indicated by the vertical dotted line in Fig. 12.22, Q0 =
0, Q1 = 1, Q2 = 1, and Q3 = 0, so n = 6. Note that in order for this simple scheme
to work, when labeling the output lines, the first is “0,” and not “1.” The circuit is
counting the pulses using base-2, also commonly referred to as “binary,” and so the
circuit is referred to as a binary counter. Various codes, including base-2 binary, will
be discussed more in the next chapter. Binary counters are available as integrated
circuit devices. Similar circuits designed to reset on the tenth pulse will count from
0 to 9 and are referred to as decade counters and are also available.

11 Alternatively, a JK flip-flop in toggle mode can be used.


258 12 Digital I

Fig. 12.21 A series of edge-triggered D flip-flops, where the output of each is connected to the
clock input of the next, can be used to create a counter circuit that runs through all the possible
output states

Fig. 12.22 The outputs for


the circuit of Fig. 12.20 as
they might be seen on an
oscilloscope

Since the binary counter with N flip-flops runs through all 2N possible combina-
tions of N 0’s and 1’s, it can be used for the scanning keyboard circuit in Fig. 12.23.12
This circuit is similar to the decoder of Fig. 12.12, except it is used for input. Clock
pulses cause the counter to run through all of the possible outputs—in this case all
of the possible row and column combinations. The decoders here are selected line
low. If the (pushbutton) switch between the row and column wire is pressed (closed)
then a low is sent to the two-input OR below. If the corresponding column line is
also low, then that OR produces a false, causing the four-input NAND to output true.
That stops the clock from getting to the counter, freezing the counter, and signals
that a key has been pressed. At that point, the output of the counter can be used as a
code that can be used to determine which key was pressed.
This type of scanning keyboard circuit, with a few extra embellishments, is what
is commonly used for modern computer keyboards. It is a priority encoder in that
whichever key is encountered first stops the process, and so pressing two keys at
once does not cause an electrical conflict. In practice, the scanning process is done
very quickly so that the keyboard response appears to be instantaneous to a human
user.

12 Any circuit that runs through all the possible outputs can be used, the binary counter often being
a convenient choice.
12.13 Other Non-logical Applications 259

Fig. 12.23 A circuit that


runs through all the possible
outputs, such as a binary
counter, can be used with
decoders to form a scanning
keyboard that checks for all
the possible key presses one
at a time

12.13 Other Non-logical Applications

All logic circuits have some propagation delay—a time between when the inputs
are set and when the output is valid. For modern digital electronics that may be
just a few nanoseconds, though can be longer for some low-power devices. While
all applications might need to consider this delay in the design, there are a few
applications that use the delay to make them work.

12.13.1 Very Short Pulse Generator

The circuit in Fig. 12.24a exploits the natural propagation delay of a logic gate to
create a small time delay between the inputs to an XOR gate. If the gates had no delay,
the output of the final XOR would always be low since the inputs would always be
equal. However, the delays cause the inputs to be unequal for a very short time, giving
rise to a very short pulsed output, perhaps just a few nanoseconds long, whenever the
input switches. The circuit of Fig. 12.24b uses external components to exaggerate
the delays between input and output and can be used for somewhat longer pulses if
one uses high input impedance (e.g., CMOS) digital gates.
260 12 Digital I

Fig. 12.24 The small delay between input and output, the propagation delay, can be used to create
short pulse from longer pulses. In a the intrinsic delay (typically measured in ns) of an XOR is used
whereas in b extra components have been added to enhance the delay

12.13.2 Oscillators

The delay between input and output can also lead to oscillations. Sometimes those
oscillations are desirable, for example to produce a clock signal. The circuit of
Fig. 12.25a uses three inverters, with feedback from the last supplying the first. Since
there are an odd number of inverters, the signal that is returned is the opposite of what
it was. Hence the circuit can oscillate due to the delay. This type of circuit is referred to
as a ring oscillator. The circuit of Fig. 12.25b uses external components to exaggerate
the delays and can make an oscillator with a frequency largely determined by the
RC time constants. Oscillator circuits, such as these, work best using logic families
that have symmetric electrical characteristics for the inputs and outputs—that is the
CMOS or NMOS families, but not the TTL families.

Fig. 12.25 The propagation delay can give rise to oscillations if an odd number of inverting circuits
is used. In a the intrinsic delay is used to get a high frequency oscillator. In b external components
are used to enhance the delay, and hence slow down the oscillations. In addition b includes an input
that can turn the oscillator on and off. These oscillators work best with CMOS logic family devices
12.14 Problems 261

12.14 Problems

1. For each of the digital logic circuits in Fig. 12.P1, use Boolean algebra to derive
the outputs in terms of the inputs. Be sure to show intermediate steps and simplify
the final result so the function of the circuit is clear. Also state the result in words
(e.g., “if all inputs are true, the output is false”).

(a) (b) (c) (d)


A A
A B
B A
S S S S
C B
B D C

Fig. 12.P1 Problem 1

2. Boolean algebra has two operators, AND and OR. A circuit that performs the
exclusive OR, XOR, signified by ⊕, was also discussed. Recall that A ⊕ B =
A · B + A · B, and use known theorems for the AND and OR operations to prove
or disprove the following:
(a) A ⊕ (B ⊕ C) = (A ⊕ B) ⊕ C (associative rule).
(b) A · (B ⊕ C) = (A · B) ⊕ (A · C) (distributive rule).
(c) A ⊕ B = A ⊕ B = A ⊕ B (see Fig. 12.1).
(d) If A ⊕ B = C then B ⊕ C = A.
3. Consider the 4-bit binary counter shown in Fig. 12.21. For that circuit, the flip-
flops are triggered on the falling edge. Describe the sequence of base-2 values
(e.g., from 12.9) that would occur if instead the flip-flops were triggered on the
rising edge, as shown in Fig. 12.P3.

Fig. 12.P3 Problem 3

4. For the circuit in Fig. 12.P4, the flip-flops are reset (0) at t = 0. Assuming a 1 kHz
clock pulse, that a logic low state is 0 V and a high state is 5 V, plot the output
of the op-amp as a function of time for t > 0. At least 8 clock pulses should be
considered (but note that 8 calculations should not be necessary).
262 12 Digital I

Fig. 12.P4 Problem 4

5. In 12.9, the Q outputs of the flip-flops are used. What happens if instead, for the
same circuit, the values of the Q outputs are used to compute n?
6. Construct a digital XOR circuit using only 2-input NORs based on the right side
of A ⊕ B = A + B + A + B.
7. Analyze the circuit of Fig. 12.P7 and describe the output as compared to the
input. Note that both logic levels and clock transitions must be included.

Fig. 12.P7 Problem 7

8. Show that the the two circuits shown in Fig. 12.P8a and b have the same truth
tables. For real circuits, what performance advantages might one of these circuits
have over the other? Write a compact truth table for these circuits.

Fig. 12.P8 Problem 8


12.14 Problems 263

Fig. 12.P9 Problem 9

9. (Challenge Problem) The circuit of Fig. 12.P9 includes three D flip-flops. The
diodes shown are LEDs arranged in a specific geometric pattern and will illu-
minate when current flows through them. The resistors keep the current from
becoming too large.
(a) Assuming the initial state has Q0 = Q1 = Q2 = 0, find the values for at least
the next six (6) clock pulses. Draw out the pattern of lit LEDs for each of
these six states.
(b) Consider the same circuit except that the central LED (“d”) is connected to
Q0 instead of Q 1 . How is the LED sequence changed?
(c) If you have solved (a) correctly, you will probably see that Q0 = Q2 = 1,
Q1 = 0 might be considered an “undesired state.” Such a state could arise
during power up, for example. How can this circuit be improved, possibly
using preset and/or clear functions (not shown), to automatically turn this
state into a desirable state (one of the six from part a) without otherwise
affecting the operation?
(Note: In the schematic, wires that cross are not connected unless there is a solid
circle present. Also assume that 0 V is the logic low, + V is the logic high, and
that + V is greater than twice the LED turn-on voltage.)
Chapter 13
Digital II

In this chapter, additional uses for logic gates are presented—in particular how the
0’s and 1’s can be used in combination as digital codes. In addition to codes useful
for numerical computation, several codes that are used for other purposes are also
presented. The process of converting between codes, including analog signals and
digital codes, is also briefly presented.
Many binary codes are shown in this chapter, most of which are presented to help
provide a broad understanding of the different ways that 0’s and 1’s can be used in
combination. Most of these schemes are not to be memorized. It is good, however,
to have some idea about what these codes are about and why they might be useful.
The codes to know in detail are base-2 binary and the related BCD code. It is also
useful to be able to read and use base-2 binary numbers written in different formats,
especially including base-2 and hexadecimal notations.

13.1 Binary and BCD Numbers

13.1.1 Binary Numbers

All codes based on a combination of 0’s and 1’s are binary. However, when referring
to numerical values, “binary” is generally used to refer to the specific case of base-2
numerical values. Base-2 values were briefly touched upon in the previous chapter
and are included here in more detail. Base-2 values are represented using a series of
0’s and 1’s, and hence are a natural numeric format for digital circuits. Rather than
the 1’s, 10’s, 100’s, etc., places used in base-10, base-2 numbers use 1’s, 2’s, 4’s, 8’s,
etc., places. Base-2 numbers are written so that the 1’s place is to the right, just as
is done with base 10. To convert from base-2 to base-10, add the appropriate powers
of 2.

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 265
B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_13
266 13 Digital II

Example Convert the base-2 value 10,110 to base 10:

10110 = 0 × 20 + 1 × 21 + 1 × 22 + 0 × 23 + 1 × 24 = 2 + 4 + 16 = 22

For a binary number, each 0 or 1, that is, each digit, is referred to as a bit. The bit
on the far right is the least significant bit (LSB or lsb), and has the smallest weight;
the digit on the far left is the most significant bit (MSB or msb), as it has the largest
weight. For the value 10,110, the LSB is 0 and the MSB is 1. Note, however, that for
the value 010,110 the MSB is 0 and not 1. For electronics and data storage, a fixed
number of bits will be used. To determine the MSB, it is necessary to know how
many bits (i.e., digits) are available.

13.1.2 BCD Numbers

In some cases, it is more convenient to use the base-10 digits we are all used to. In
order to represent those digits using 0’s and 1’s with digital circuits, some sort of
alternate binary code is used for the digits. Any such code that uses 0’s and 1’s to
represent base-10 digits is a form of “binary coded decimal” or BCD. A minimum
of four-bits is necessary to represent the digits 0–9. In that case six of the possible
four-bit codes are not used.
Normally, when referring to BCD, the scheme is to represent each digit with its
base-2 representation, and then those four-bit digits are used as the digits of a base-10
value. That is, instead of using the normal base-10 symbols, each digit is represented
by a sequence of four bits, which is also the base-2 representation of that digit.
Example Convert 375 decimal (base 10) to BCD.

3 = 0011; 7 = 0111; 5 = 0101


375 → (0011)(0111)(0101) = 001101110101

In the same way that, when referring to numerical values, “binary” is generally
used to mean base-2, BCD without any other qualifiers usually refers to this use of
the base-2 representation of the decimal digits. There are many other ways to use
0’s and 1’s to represent the digits 0 to 9 that are also forms of BCD. Some examples
will be illustrated later.
13.2 Other Weighted Binary Codes 267

13.1.3 Hexadecimal and Octal Notation

While electronics works well using base-2 binary, people like to use decimal (base-
10). As a compromise, base-8 (“octal”) or base-16 (“hexadecimal”) notation is some-
times used to write base-2 binary numbers in a more compact way. Each of these is,
in some sense, “closer” to base 10 than is base-2.
In octal, the digits 0–7 are used and each octal digit represents three base-2 digits.
The digits 8 and 9 are illegal. In hexadecimal, there are 16 digits: the digits 0–9 are
used plus A, B, C, D, E, and F for the digits corresponding to the base-10 values
10, 11, 12, 13, 14, and 15 respectively. Each hexadecimal digit corresponds to four
base-2 digits. For octal and hexadecimal there are no unused binary codes.
Example Convert the base 2 value 10,110,101 to octal and hexadecimal.
• Octal: put bits in groups of three starting on the right (add zeros on the left as
necessary), and write the corresponding digit.

(010)(110)(101) = 265.

• Hexadecimal: put bits in groups of four starting on the right (add zeros on the left
as necessary), and write the corresponding digits, using A through F for the digits
corresponding to 10–15.

(1011)(0101) = B5.

It is much more common to see hexadecimal than it is to see octal. To distinguish


hexadecimal values from decimal (base-10) values, an additional character is some-
times included. For example, 123 h, $123, #123, and 0 × 123 are some of the many
notations that are used to indicate that the value is to be interpreted as hexadecimal.
To convert to decimal, multiply each digit by the appropriate power of the base,
starting on the right. Remember to start the powers with 0 and not 1.
Example Convert 1B3h to base-10:

3 × 160 + 11 × 161 + 1 × 162 = 3 + 176 + 256 = 435.

13.2 Other Weighted Binary Codes

A weighted binary code consists of N “bit” positions and each position is assigned a
numerical weight. The base-2 binary code is an obvious example where the weights
are (from right to left) increasing powers of two. Binary Coded Decimal (BCD) is a
mixed weighted code. For each group of four bits, BCD is identical to base-2 binary.
If the bit position contains a 1, that weight is added to the result, if the position holds
268 13 Digital II

a 0 that weight does not contribute. Then each group of four bits has a weight which
is a power of ten.
There are other weighted codes that are sometimes useful.

13.2.1 The 4221 Code

Consider a four-bit code with weights 4, 2, 2, and 1. This code is to be used to


represent the decimal digits 0–9. It is clear that no matter what selection of bits is
chosen, the result will never be outside the range from 0 to 9. Some digits can be
represented in more than one way. In the 4221 code, the ambiguity is resolved using
the following rules.
• For the digit 5, use 2 + 2 + 1 instead of 4 + 1.
• If the digit is less than 5, use the rightmost 2 first, if needed.
• If the digit is greater than 5, use the leftmost 2 first, if needed.
The resulting codes for the decimal digits are shown in Table 13.1.
Such a code could also be considered a form of BCD, though in this case “binary”
does not refer to base-2, but simply that there are two values (0 and 1). Again, some
possible codes are illegal and are not used. This code has the property that for any digit
N, the code for 9-N (known as the “9’s complement”) is also the one’s complement
of N. That is, to get the 9’s complement, change all zeros to ones and all ones to zeros,
a task that is easily accomplished in the electronics with NOT gates (inverters). For a
multi-digit number, do this for each digit. This makes the 4221 code convenient for
implementing some decimal arithmetic using electronics (via the 1’s complement, or
NOT gate).1 Another code that has this same property is the “excess 3” code, where
each digit is represented by its binary value plus three (e.g., the digits 4 = 0111, 5
= 1000, etc.). The excess 3 code is not a weighted code due to the offset by 3, but is
close.

13.2.2 2 of 5 Codes

Consider a five-bit code to be used to represent the digits 0 to 9 only. A code can be
constructed where exactly two of the five bits are 1’s, the remaining being zero. There
are exactly ten unique ways to select two things out of five. This gives 10 unique
codes that can represent the ten digits used for numerical values. The assignment

1 To subtract Y from X, add the 9’s complement of X to Y, then take the 9’s complement of the
result. If X = 492, then its 9’s complement is 507. Hence 492–237 = 9’s complement of (507 +
237 = 744) = 255. The computation can also be accomplished using the 10’s complement of Y,
which is its 9’s complement plus 1. Add the 10’s complement of Y to X, then discard any extra
digits generated on the left. So 492–237 → 492 + (762 + 1) = 1255. Discard the extra “1” on the
left to get 255. In either case, the subtraction has been changed to addition.
13.2 Other Weighted Binary Codes 269

Table 13.1 4221 codes for


Digit Code Digit Code
the digits 0 to 9
0 0000 5 0111
1 0001 6 1100
2 0010 7 1101
3 0011 8 1110
4 1000 9 1111

of which code corresponds to which digit can be done in an arbitrarily, however it


is usually done in a way that can be interpreted as a weighted code. For example,
the weights 7, 4, 2, 1, and 0 can be used. In this scheme each code corresponds to
a digit obtained by adding the appropriate weights for the two locations where the
bits are 1’s. The exception is the digit “0,” that is assigned the one remaining code
corresponding to weights 7 + 4 = 11. The weights 6, 3, 2, 1, and 0 could be used
instead. The weights may be used in any order—largest to smallest and smallest to
largest being the most common.
For multiple digit numbers, simply run the codes together. It is hoped that the code
is unique enough that any stray signals can easily be discarded as “illegal codes.”
Example For 7, 4, 2, 1, 0 weighting, decode “000,111,100,010,010”.
Put in groups of 5 − (00,011)(11,000)(10,010), and apply the weights to get “108.”
For this code, a sequence of five 1’s in a row will never occur. That fact can be
used to create a special code of five 1’s in a row that can be used to indicate the
beginning and/or the end of the data transmission. While not always utilized, it also
provides a means for the receiver of the code to be synchronized to the transmitted
data before the data starts.
Note that, in principle, a 5-bit code can represent 32 different items. Like other
BCD codes, the 2 of 5 codes do not use all the possibilities. This can be useful to
detect errors. For the 2 of 5 code most errors result in an illegal code that is easily
detected. Other similar codes exist, though they may not be weighted. For example, a
3 of 6 code has 20 possible unique codes making it suitable for sending hexadecimal
digits (with a few leftover), and is “balanced” in that there are the same number of
zeros as ones. Balanced codes have an advantage for some transmission methods.
Some more advanced coding schemes will encode several digits at once, rather
than one at a time. For example, “5 of 13” and “8 of 13” codes are used as one part
of some current postal bar code schemes. There are over 1000 possible unique codes
for each.
270 13 Digital II

Table 13.2 Decimal to gray


Value Value
code representation
Decimal Gray code Decimal Gray code
0 0000 8 1100
1 0001 9 1101
2 0011 10 1111
3 0010 11 1110
4 0110 12 1010
5 0111 13 1011
6 0101 14 1001
7 0100 15 1000

13.3 Non-weighted Codes

13.3.1 Gray Code

The Gray code2 is a simple example of a non-weighted code where the conversion
to and from base-2 binary (a weighted code) is straightforward. The Gray Code is
designed so that only one bit changes from one number to the next. One way to
express the basic rule for counting up from zero is that for each new value, change
the lowest (rightmost) bit that results in a new code. Counting from 0 using a (four
bit) gray code would generate the values in Table 13.2.
Such a code is useful for some analog to digital conversions so that noisy data
near a boundary between two values will result in one or the other of two adjacent
values. An example of such a system is to encode a rotating mechanical position
using a Gray code wheel. One possible 4-bit wheel is illustrated in Fig. 13.1a. The
wheel might be installed on a shaft with sensors such as shown in a side view in
Fig. 13.1b. Here the white areas in the disk would be clear and the dark areas opaque
(or vice versa). If a base-2 binary code is used, accurate results would require that
multiple sensors switch simultaneously at the boundaries. For example, to go from
the number 7 (0111) to the number 8 (1000) in base-2 requires all four bits to switch
simultaneously. If they do not switch at exactly the same time, some other value,
possibly very far away from both 7 and 8, will be present during the transition. If
Gray code is used, then only one bit changes and the result at that boundary will be
either 7 or 8, and will never venture far away.

2 Named after Frank Gray (1887–1969) who patented a coding method using the code in 1947. He
referred to the code as a “reflected binary code.” Gray codes are also an important part of Karnaugh
maps, that are sometimes used to simplify digital circuitry. There are now a number of variants of
the original code that is described here.
13.4 Bar Codes 271

Fig. 13.1 (a) An example of one type of Gray code wheel and (b) how such a wheel might be used

13.3.2 The ASCII Code

Other non-weighted codes are simply assignments between certain codes and certain
“values” that may not even be numeric. One such example is the ASCII code, which
is by far the most common way to represent letters and characters using 1’s and 0’s.
The original standard US ASCII code uses 7 bits. Modern computers extend most
character sets to 8 bits or more. Tables of ASCII codes for different alphabets (or
symbols) are easy to find, and can be used to look up the code for each character.
The original ASCII code included many control functions. Many of these control
functions are now obsolete and only a few will actually appear in practice. In many
applications, the control codes are now replaced with printable characters. Initially
these codes were to be produced by holding down the control (Ctrl) key on the
keyboard while typing the corresponding letter.3 Some of these control codes have
evolved to behave differently depending upon context. For example, on many systems
the Ctrl-G code (“bell”) will ring a bell only when it is being “displayed” and not
when it is being “entered” and while Ctrl-I used to be the same as “Tab,” in some
systems it is now used to toggle italic fonts on and off.

13.4 Bar Codes

For binary data transmission it is common to send and receive serial data—data
sent one bit at a time—sometimes at an arbitrary rate and sometimes in “noisy”
environments. Since the receiver does not necessarily know the speed of transmission,
or even if a transmission has occurred, the receiver must be able to figure that out
from what is received. The following are illustrations of codes designed to help do
this. What are shown are examples of bar codes, which are easily examined visually.

3 This is where the “Ctrl” key comes from. To generate the control codes, one would hold down the
Ctrl key and press the appropriate letter. For example, Ctrl-H is a backspace, Ctrl-M is a carriage
return, Ctrl-[ is the Esc key, etc.
272 13 Digital II

Fig. 13.2 An example of an


interleaved 2 of 5 bar code

The same practices can be used for binary data sent electronically between devices
over cables or via radio signals.

13.4.1 Interleaved 2 of 5

A bar code can be implemented using fat bars and thin bars to represent 1’s and 0’s.
Noting that the white space between the bars can also be fat and thin, this gives rise
to the “interleaved 2 of 5 code” where each group of 5 bars represents a digit, where
two of the bars are fat and three thin, and the 5 spaces represent a second digit where
two of the spaces are fat and three are thin. A start code (commonly a thin bar-thin
space-thin bar-thin space) and a stop code (commonly a fat bar-thin space-thin bar)
are also defined, as is a minimum “quiet zone” near the bar code, where no bars or
other markings are allowed. The differing start and stop codes allow the code to be
read from either end and still be decoded correctly. An example of an interleaved 2
of 5 bar code is shown in Fig. 13.2.
Codes such as, and similar to, the interleaved 2 of 5 code can observed on many
cardboard shipping cartons.

13.4.2 UPC Codes

The UPC codes used for identifying products in stores are somewhat more compli-
cated than the 2 of 5 code, though are still forms of BCD. Only the 12 digit “UPC -
A” is discussed here. It has the following properties for coding of digits:
• The code for each decimal digit is a sequence of 7 bits.
• As you traverse the 7 bits, the bits change exactly 4 times—the first change being
the first bit, that by definition changes from whatever preceded it. There are 20
such codes—only 10 if one also considers codes in pairs.
• Whether you start with a “1” or “0” is irrelevant, only the changes matter.
The codes for the digits in UPC written in terms of the number of like bits in a row
are shown in Table 13.3. There is no simple relationship between the code and the
13.4 Bar Codes 273

digit. The codes were chosen in a manner that was thought to reduce the possibility
of read errors during normal use.
Note that the codes in reverse are also unique and correspond to the same digit –
digits will not be confused when the code is read backward. When this code is turned
into a bar code, the values above for the condensed code correspond to the width of
bars and spaces. Each digit is 7 units across, with two bars and two spaces. Figure 13.3
illustrates an example, showing the two codes for the digit 5.
The UPC code also has special codes at the beginning, middle and end, referred to
as “guard bars.” These bars also make it so that the six digits on the left will always
start with a space while the six on the right start with a bar.
The first digit of the UPC code is used to describe the type of product. The next
5 digits are a manufacturer’s code, the five after that are a product code and the last
digit is a check digit. To compute the check digit, add the digits in the 2nd, 4th, etc.,
positions, plus three times the digits in the 1st, 3rd, 5th, etc., positions. Then the
smallest integer that can be added to get to a multiple of ten is the check digit. When
scanning, the check digit helps to ensure an accurate scan. An example of a UPC
code is shown in Fig. 13.4. Similar ideas are used to send “packets” of information,
which can contain much more information, over cables or using radio signals. The
decimal digits across the bottom of the UPC code are for reading by a human, in the
event that automatic scanning does not work.
Many binary code transmission protocols are also based on the use of changes to
indicate bit values. That is, a voltage level of 0 V or 5 V is not important, but when

Table 13.3 7-bit codes used


Digit Code (condensed) 7 bit codes
to represent decimal digits for
UPC symbols 0 3211 1,110,010 or 0,001,101
1 2221 1,100,110 or 0,011,001
2 2122 1,101,100 or 0,010,011
3 1411 1,000,010 or 0,111,101
4 1132 1,011,100 or 0,100,011
5 1231 1,001,110 or 0,110,001
6 1114 1,010,000 or 0,101,111
7 1312 1,000,100 or 0,111,011
8 1213 1,001,000 or 0,110,111
9 3112 1,110,100 or 0,001,011

Fig. 13.3 Examples of the


two ways to represent the
digit 5 in a UPC code
274 13 Digital II

Fig. 13.4 An example UPC


code

the voltage level either changes, or does not change, from one to the other, that is
where the information is contained.
Fancier “bar codes” include the 2-dimensional codes, such as “QR codes,” which
are not discussed here. Those codes are much more sophisticated, can represent both
numbers and characters, and include error correcting in that even a damaged code
can be correctly decoded.

13.5 Some Numeric Code Conversions

When codes are developed systematically, rather than by random assignment, it


is usually possible to convert from one to another relatively simply. Of particular
interest here are conversions that are easy to do within digital electronics, and not
necessarily those that are done using a human with a hand calculator. In what follows,
to “complement” refers to the ones complement—change all of its 0’s to 1 and 1’s to
0. Also recall that LSB and MSB refer to the least significant (right-most) and most
significant (left-most) bits of a stored value with a fixed number of bits.

13.5.1 Base-2 Binary to Gray Code

A procedure to convert from base-2 binary to Gray code is as follows:


Initially, set X = binary number (integer) to convert and Y = 0.
• Shift X and Y left by one bit; MSB of X → LSB of Y
• If LSB of Y = 1, then complement X
• Repeat 1 and 2 for all bits of X
Y will now contain the result.
13.5 Some Numeric Code Conversions 275

Fig. 13.5 Circuits to (a) convert base-2 binary to Gray code and (b) convert Gray code to base-2
binary

13.5.2 Gray Code to Base-2 Binary

One way to convert from Gray code to base-2 binary is as follows:


Initially set Y = Gray Code Number (integer) to convert and X = 0.
• Shift Right both X and Y by one bit; LSB of Y → MSB of X.
• if LSB of Y = 1, complement X.
• Repeat steps i and ii for all bits of Y.
X Will Now Contain the Result.
The two conversions between base-2 binary and Gray code can be implemented
with electronics as shown in Fig. 13.5 (shown for 4 bit values, but extends to any
length).

13.5.3 Decimal to Gray Code

While a conversion from decimal to Gray code could be accomplished by first


converting to base-2 binary, and is probably faster when done that way, the conver-
sion can be done directly. Since Gray code is not a weighted code, it is not obvious
that such a conversion should be possible.
The procedure is as follows:
Set Y equal to the positive decimal integer to be converted into an N-bit Gray
code value (2N > Y ).
Fill in the N bit positions of the result from the left as follows:
• Add 1/2 to Y (this takes care of rounding issues).
• If Y ≥ 2 N −1 then the MSB is 1, otherwise it is 0.
• Set m = N − 1.
• Set X = |Y − 2m |
• if X ≤ 2m−1 the m-th bit is a 1, otherwise it is 0.
276 13 Digital II

• Set Y = X, and decrement m by 1.


• Go back to step 4 and repeat until m = 1.
Note that the first comparison, in step 2, a greater than or equal sign is used,
whereas all later comparisons, e.g., in step 5, it is a less than or equal to sign. Because
of the added ½ in step 1, the “equal” part of those signs is somewhat superfluous.
Example Convert 37 base-10 to 6-bit Gray Code.
• 37 + 0.5 = 37.5
• 37.5
| ≥ 25 , |so 6th bit is 1.
• |37.5 − 25 | = 5.5
• | ≤ 2 4, |so 5th bit is 1.
5.5 4

• |5.5 − 2 | = 10.5
• 10.5
| ≥ 23 , |so 4th bit is 0.
• |10.5 − 23 | = 2.5
• | ≤ 2 2, |so 3rd bit is 1.
2.5 2

• |2.5 − 2 | = 1.5
• | ≤ 2 1, |so 2nd bit is 1.
1.5 1

• |1.5 − 2 | = 0.5
• 0.5 ≤ 20 , so 1st bit is 1.

The result is that 37 base-10 is 110111 in Gray code.


A circuit to implement this is shown later as part of the discussion of analog to
digital conversion.

13.5.4 BCD to Binary Conversion

Here BCD is used in its usual sense to mean “base-2 binary coded decimal.” Of
interest here is a direct conversion from BCD to binary, without an intermediate
conversion to decimal. That is, how it would be done with electronic circuits.
Initially set X = the (integer) value to convert (written as a string of BCD digits)
and set Y = 0.
(1) Shift X right by one bit, with LSB of X = first (next) bit of binary number, Y
(2) Examine each new 4-bit digit of X. If a digit is greater than 7, subtract 3 from it.
(3) Repeat steps 1 and 2 until X contains only zeros.
13.5 Some Numeric Code Conversions 277

Y now contains the result.


Example Convert 49 to base-2 binary.
X = 49, written as a two-digit BCD number = (0100)(1001), (Parentheses added to
show groups of 4) and Y = 00,000,000 initially.

Thus, the BCD value 01,001,001 equals the base-2 value 00,110,001.

13.5.5 Binary to BCD Conversion

To go from binary to BCD, run the above algorithm in reverse. Let Y = binary number
to convert, X = 0 initially.
(1) Examine each group of four bits of X, add 3 to all that are greater than 4 (this is
trivial the first time through, since they are all 0).
(2) Shift X and Y left one bit, 0 → LSB of Y, MSB of Y → LSB of X.
(3) Repeat steps 1 and 2 until Y = 0.
X Now Contains the Result.
Example Convert Y = 110,001 in (base-2) binary to BCD.
278 13 Digital II

X is now the BCD representation of 49 (0100 = 4, 1001 = 9).

13.6 Digital to Analog Conversion

It is often desirable to convert stored digital information into a continuous analog


signal (i.e., a non-digital signal). For example, music stored on a digital device (a
CD, an MP3 player, a downloaded data stream, etc.) is stored as a series of zeros and
ones, however the sound those numbers represent, and the sound you want to listen
to, is a continuous analog signal. The conversion of a digital value to an analog value
is referred to as digital to analog, or D/A (“D to A”), conversion.

13.6.1 The 1-Bit D/A

A very simple and inexpensive digital to analog D/A scheme uses just 1-bit for
output. This scheme works as long as the maximum frequency components within
the signal are quite low in frequency compared to the maximum frequency of the
digital electronics. Music and voice recordings are prime examples where this can
be true.
The simplest D/A of this sort is based on the duty cycle. During an interval, T, the
output is a logic 1 for a fraction of the time proportional to the desired output, and
off the rest of the time. The average during the time T represents the analog value.
13.6 Digital to Analog Conversion 279

Fig. 13.6 Examples of (a) pulse width encoding and (b) pulse density encoding. In each case the
signal is sent through a low-pass filter to extract the average value

This simple scheme is called “pulse length encoding” or “pulse width encoding” and
is illustrated in Fig. 13.6a. The output is then sent through a low-pass filter resulting
in the time averaged signal. If the highest desired frequency is f max , then it must be
the case that 1/T >> f max and the filter would be designed to have a low-pass cut-off
to block frequencies higher than f max .
A second similar scheme uses pulses of constant duration, but the spacing between
the pulses is variable. Once again, the analog value is represented by the duty cycle
and is extracted by sending the pulses through a low-pass filter. Such a scheme would
be referred to as “pulse density encoding” and is illustrated in Fig. 13.6b.
The so-called delta-sigma 1-bit D/A (and also A/D) is a variant of this simple
scheme that switches the output on and off many times during the time interval T.
In this case, the values are related to the time rate of change (the slope) of the signal,
and to convert to analog, the sum of the changes (a numerical integration) is used.
Consider two people traveling along a sidewalk. One of them can walk forward
and/or backward continuously at any speed. The other, perhaps due to a childhood
curse, can only hop by 1 m each second, though the hop can be either forward or
backward. To travel together, the second will hop forward if behind the first, and
hop backward if ahead. Clearly, if the first walks faster than 1 m/s, then the second
will not be able to keep up. However, as long as the first is not traveling too fast, the
second will always be within 1 m. If a record is kept of the direction of each hop,
the approximate location, within 1 m, of the first person can be reproduced from the
recorded directional values. To reconstruct the values, that is to convert the digital
data back to analog, add the number of forward hops and subtract the number of
reverse hops. This scheme is illustrated in Fig. 13.7. In Fig. 13.7a, the signal and the
digital approximation are shown. Figure 13.7b, shows the stored data, that indicates
the direction for each successive hop (a “1” is forward and a “0” is reverse).
The real advantage of all of these 1-bit schemes is that only digital logic gates are
required—no linear circuits other than a simple low-pass filter are necessary. In fact,
for audio signals the mechanical properties of a speaker or earphones are often used
to provide much of the low pass filtering required.
280 13 Digital II

Fig. 13.7 At (a) a signal (solid line) is approximated with a “hopping” digital value. The direction
of each hop is stored, as shown in (b)

13.6.2 A Summing D/A

A simple D/A that can respond at higher speeds uses a summing op-amp circuit.
The resistors are chosen to provide the appropriate binary weights to the signals.
Using appropriate resistors, such a circuit can be used with any binary weighted
code. Figure 13.8 shows an example using a binary base-2 code with four bits.
Due to the wide range of resistance values required, requiring a very high level
of precision for the larger resistors, these circuits are practical only when there are a
relatively small number of bits.
An improved circuit for binary base-2, that expands nicely to as many as 12 or
even 14 bits, uses a series of identical resistors in a “ladder” circuit such as is shown
in Fig. 13.9 (also see the R-2R ladder example in Chap. 2). To get a good match, two
resistors with resistance R in series can be used to get an equivalent with resistance
2R, or two “2R” resistors in parallel to get R. Using this trick, all the resistors used
can have the same value. With current technology, equal-valued resistors can be
matched to each other very precisely. In addition, rather than using the logic levels as

Fig. 13.8 A simple


summing D/A converter for a
base-2 binary weighted code
13.7 Analog to Digital Conversion 281

Fig. 13.9 An “R-2R ladder”


approach to convert base-2
binary values to analog
values

voltage sources, the logic levels are used to control (transistor) switches connected
to a highly regulated reference voltage in order to obtain more controlled and precise
values. Such D/A devices are available as a single integrated circuit.

13.7 Analog to Digital Conversion

When collecting data from a measurement, digital codes generated and stored should
represent the corresponding measured values (i.e., voltages or currents). For digital
storage, the first step is to convert from the non-digital (analog) signal to a digital
signal. This is called analog to digital conversion or A/D (“A to D”). Digital data can
only have discrete values and what one really wants to do is to obtain the best digital
approximation to the continuous analog data. The Gray code wheel already discussed
above is a form of A/D that can provide, for example, a digital representation of the
continuous angle of the rotation of the wheel.
A few schemes are discussed below related primarily to conversion of analog
signals (e.g., voltages) to binary base-2 values. Some of these schemes can be adapted
to other codes as well.

13.7.1 Voltage to Frequency Conversion

A particularly simple, though slow, method to convert a voltage to a digital value


is to use a voltage-controlled oscillator (VCO). A VCO produces a periodic signal
with a frequency that depends on an input voltage. VCOs are available as integrated
circuit devices. The frequency from the VCO is measured to obtain a digital value.
That is, output pulses from the VCO are sent to a counter circuit for a fixed time
interval. For measurements of very slowly changing values, such measurements can
be very inexpensive and may have very high precision.
282 13 Digital II

Fig. 13.10 Timing how long


it takes a voltage ramp to
match an input signal is an
inexpensive A/D method.
The ramp is constructed with
a constant current source and
a capacitor

13.7.2 Timing Schemes

Another simple and inexpensive, though relatively slow scheme, is to generate a


sequence of voltages using a linear ramp, and then count how long it takes for the
generated voltage to pass the input voltage. To obtain a linear ramp, a constant
current source is used to charge a capacitor, as shown in Fig. 13.10. While the
capacitor is charging, clock signals are sent to a counter to measure the elapsed time.
A comparator detects when the ramp passes the input voltage and then stops the
counter. The digitized result is taken from the outputs of the counter.
Note that for this conversion scheme it is also a simple matter to obtain results in
BCD, such as is appropriate for a simple digital voltmeter. Readily available BCD
counters are used instead of binary counters. Also note that the switch to ground that
discharges the capacitor, shown as a manual switch, is normally implemented with
a transistor and some control circuitry to start and stop the measurement.

13.7.3 Search Schemes

In a search scheme, a digital guess is made, converted to analog using a D/A, and
then compared to the input. Based on the result, a new guess is formed until the
desired accuracy is achieved. This is illustrated in Fig. 13.11.
The simplest search might start with zero and count up, which is almost identical
to the timing scheme mentioned above. It is much more efficient to start with the most
significant bit and work backward. To start with zero and count up is analogous to
finding an entry in the dictionary by starting at the beginning and reading each entry
until a match is found. When starting with the most significant digit, the dictionary is
opened to the middle and it is determined whether that was too far or not far enough.
Based on that result, proceed forward or backward half of the remaining distance to
the beginning or end. Then based on the result there, go half of what is left toward
13.7 Analog to Digital Conversion 283

Fig. 13.11 A searching A/D provides a digital guess and then refines that guess based on the results

the beginning or end. If there are 2N entries, this latter scheme will take N searches,
whereas the former would take, on average 2N −1 searches. If N is large, the difference
can be quite significant.4
Circuits that do this type of conversion with good accuracy are available as
integrated circuit devices.

13.7.4 Analog to Gray Code Conversion

Converting directly to Gray code, a non-weighted code, is more difficult to set up,
but can result in a very fast conversion. Like the simple summing D/A, the precision
of components will limit the number of bits to a half dozen or fewer. This scheme
shown here was developed by [1]. The method is a literal electronic implementation
of the decimal to Gray code conversion scheme seen earlier and the circuit is shown
in Fig. 13.12.
In Fig. 13.12, the differential amplifiers have a gain of 1. The absolute value
circuits can be any of those described previously (see Chap. 10 and the beginning of
11). The reference voltages are given by
( )
V0 N −1 1 V0 (N −m)
VN −1 = 2 − ; VN −m = 2 , (13.1)
2N 2 2N

where the input range is 0 to V 0 , and 1 < m ≤ N. Subtracting ½ for the first value is
equivalent to adding ½ to corresponding input value, which is done to take care of
rounding errors. The Gray code output is taken from the comparators.

4There are about 200,000 words in a comprehensive English dictionary. Thus, N = 18 and any
word can be found with no more than 18 of these operations.
284 13 Digital II

Fig. 13.12 A circuit to convert analog levels directly to Gray code

13.8 Quantization Noise

With N digital bits of information, only 2N discrete values can be represented


precisely. Thus, the best any analog to digital converter can do will be to find which
of the 2N values is closest to the analog value. The possible digital values are discrete
and quantized. The difference between the analog value and the closest digital value
is referred to as quantization error or quantization noise. For example, with 7 bits,
128 distinct values can be produced. If these values are spread evenly through the
measurement range, then the accuracy that any digital result will likely match an
analog signal is no better than 1 part in 128, or about 1%, of the full measurement
range.
For many types of measurements, the measurement procedure can be repeated
multiple times with the expectation that the average of many results will more
precisely represent the “true” value. With a low-noise signal and quantization error,
that may not happen. If the variations in the signal are smaller than the spacing
between the quantized levels, the same error can occur with every measurement—it
is not a random error and will not average to zero. In fact, in such situations, one
strategy for improving the accuracy is to add random noise to the signal. If the added
noise averages to zero and is larger than the quantization spacing, then the average of
many measurements can, in fact, be more accurate than measurements made without
the additional noise.

13.9 Problems

1. Convert the decimal number 1234 into base 2, octal, and hexadecimal.
2. Convert the base-2 binary value 001,001,001,001 into decimal.
3. Convert the BCD value 001,001,001,001 into decimal.
4. Convert the decimal value 1234 into BCD.
Reference 285

5. The nine’s complement of a digit N is 9 − N, where 0 ≤ N ≤ 9. For a multi-digit


value, the nine’s complement is formed by taking the nine’s complement of each
digit. The ten’s complement of a decimal value is its nine’s complement plus 1.
(a) Compute 573 minus 250 using your usual method.
(b) Compute 573 plus the (three-digit) ten’s complement of 250, and then keep
only the lowest 3 digits in the result.
(c) Compute the sum of the nine’s complement of 573 with 250, then take the
nine’s complement of the result.
6. Web pages often use hexadecimal values to define colors. The values represent 24
bits. The 24 bits are three groups of 8 bits to specify, in order, the amount of red,
green, and blue that contribute to the color. Thus, in decimal, each of the three
contributions can range from a minimum of 0 to a maximum of 255. Suppose a
color is specified as “ffc033.” What are the weights, expressed in decimal, for
the three contributing colors?
7. If an A/D produces a 6-bit result, providing evenly spaced values representing
values for the range from 0 to 5 V, what is the spacing, in volts, between the
evenly spaced values?
8. If analog values in the range between 0 and 10 V are to be recorded and stored
using an N-bit digital value, what is the minimum value of N that can be used if
an accuracy no worse than 0.1% is desired?

Reference

1. Y.K. Yuen, Analog-to-gray code conversion. IEEE Trans. Comput. 27, 971–973 (1978)
Chapter 14
Calculators and Computers

Calculators, computers, microprocessors, and the like are all based on digital
circuitry. This chapter discusses how digital circuits can be combined to perform
numerical calculations and the basics of how computer processing is accomplished.
Along the way, tri-state devices are introduced as well as some unique additional
applications that they enable.

14.1 Adding Base-2 Numbers

The truth table for the addition of two one-bit values, A and B, is shown in Table
14.1, corresponding to 0 + 0 = 0, 0 + 1 = 1 + 0 = 1, and in base-2, 1 + 1 = 10. A
simple circuit to add two bits is shown in Fig. 14.1a. Such a circuit is referred to as
a “half adder.”
For multiple-bit values, however, provisions for a possible carry from a previous
operation need to be included—there will be three bits to add. That is accomplished
using a “full adder.” A full adder can be constructed from two half adders as shown in
Fig. 14.1b. To add two N-bit values requires N full adders, as illustrated in Fig. 14.2
for N = 4. The result of the addition, S, also has N bits. If the final carry is 1, then the
final carry is lost from S and S is said to have “wrapped around” zero. This behavior
can be a problem but can also be used to advantage for subtraction using the two’s
complement method.

14.2 Two’s Complement Arithmetic

Subtraction using binary numbers can be accomplished using a subtraction circuit,


similar to the full adder just shown. More commonly, however, subtraction is done
as an addition using two’s complement values—a process somewhat similar to, but

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 287
B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_14
288 14 Calculators and Computers

Table 14.1 Truth table for 1


A, B Sum
bit addition
0, 0 0
0, 1 1
1, 0 1
1, 1 0, carry a 1

Fig. 14.1 Circuits for (a) half adder and (b) full adder

Fig. 14.2 An adder for


four-bit values, constructed
from four full adders

not the same as, performing subtraction by the addition of a negative value. With the
two’s complement method there is no need for a separate subtraction circuit.
When a digital circuit is used to add two numbers, the circuitry will always add a
fixed number of binary digits (0s and 1s)—that is, a fixed number of bits. This will
be true no matter how many bits are actually required to represent the values being
added. Often, binary values are used in the electronics (the “hardware”) in groups
of 8 bits, known as “bytes,” 16 bits, which are called “words,” 32 bits, which are
14.2 Two’s Complement Arithmetic 289

“double words” or 64 bits, sometimes referred to as a “quadword.”1 That there is a


fixed and finite number of bits is a limitation, but also an opportunity.
If there are N bits, each of which can be a 0 or a 1, they can represent 2N binary
numbers ranging from 0 to (2N − 1). Alternatively, one of the bits can be used to be
a minus sign and then there are 2N values from −(2N −1 ) to +(2N −1 − 1). Rather than
a simple minus sign, there is another way to represent negative values for simple
integers that is called the method of two’s complements. The two’s complement
method is only effective when there are a fixed number of bits.
To form a two’s complement of a base-2 number, do the following two steps:
• Make the one’s complement of each binary digit by changing all 0’s to 1’s and all
1’s to 0’s.
• Add 1 using normal binary addition, throwing away any carry from the MSB.

Example Find the two’s complement of 0101100.

• Make the one’s complement: 1,010,011.


• Add 1: 1010011 + 1 = 1010100.

Subtraction is then accomplished by adding the two’s complement.

Example Compute 67 − 13 for 8-bit binary integers.

• Convert decimal values to 8-bit binary:

67 = 64 + 2 + 1 = 26 + 21 + 20 = 01000011
13 = 8 + 4 + 1 = 23 + 22 + 20 = 00001101

• Form the 2’s Complement for 13, the “subtrahend”:

1' s Complement = 11110010


Add 1 = 11110010 + 1 = 11110011

• Add the two values, keeping only the lowest 8 bits (the “1” in parentheses is
thrown away, there is no place for it).

1 1 1

01000011
 11110011
(1)00110110

1 Though rarely seen in hardware any more, a group of 4 bits, being half of a byte, is referred to as
a nibble.
290 14 Calculators and Computers

• Convert back to decimal to check:

21 + 22 + 24 + 25 = 2 + 4 + 16 + 32 = 54

Two’s complement arithmetic is a “clock” arithmetic. For example, on a twelve-


hour clock 1 o’clock minus 2 h is 11 o’clock. Hence, one can also interpret 11 o’clock
as “ −1 o’clock” since 1 − 2 = −1. A negative value expressed as a two’s complement
will always have a “1” in the MSB, however that is not simply a minus sign. For
example, for 8-bit two’s complements, consider the values, −1 = 11111111, −2 =
11111110, etc. The initial “1” signals that the value is negative, but the remaining
binary number is not the magnitude of that value.

14.3 A Simple Arithmetic Logic Unit (ALU)

Rather than create separate circuitry to perform each operation, many potential oper-
ations can be built into one circuit. That circuit has additional inputs that tell it
which operation to apply this time. One example of such a circuit is an Arithmetic
Logic Unit (ALU). An ALU performs the basic functions for computations and logic
operations.
Figure 14.3 shows a simple example of such a circuit that includes 4 “command”
input bits that determine what operation is to be applied to the bits of the two input
values (A and B). The outputs include a value S along with some “flag” bits. The flag
bits indicate general features of the result. Here A is an n-bit binary (base-2) value
represented by the bits Ai , i = 0 to (n − 1), and similarly for B, S, a, and b. The
single-bit flag values “c” and “z” will depend on the values of the input arguments
and commands. In the diagram, a “register” is (anything equivalent to) a set of D
flip-flops that store values after a clock pulse. A description of what the ALU circuit
does can be presented as an abbreviated truth table, such as Table 14.2. Notice how
easy it is to add the two’s complement of B to A, a subtraction, through the use of
the command bits and the full adder.
Inside a (simple) microprocessor, it is common to find the ALU output register
directly wired back to one of the inputs, such as the A input. That A input will then
be referred to as the accumulator. That is, the output from any operation is wired to
(potentially) become one of the inputs for the next operation. If it is assumed that
the output bits (S0 to Sn−1 ) are indeed connected directly to the A input bits (A0 to
An−1 respectively), and the B inputs (B0 to Bn−1 ) are coming from a register called
“B,” then the sixteen commands for this simple ALU can be described as shown in
Table 14.3.
Of course, in a real microprocessor (or similar device) many additional commands
would be expected. Additional operations would include shifts, various logic opera-
tions (AND, OR, XOR, etc.), and others. This ALU is only a simplified example to
illustrate the principle that is used.
14.3 A Simple Arithmetic Logic Unit (ALU) 291

Fig. 14.3 A simplified arithmetic logic unit (ALU) that operates on two input values, A and B, and
performs an operation determined by the command values C

Table 14.2 Summary of the functions of the circuit in Fig. 14.3


Command input bits Full adder Outputs Description
inputs
C3 C2 C1 C0 A b S Carry Zero
0 0 0 0 0 0 0 0 1 Clears to zero
0 0 0 1 0 0 1 0 0 Sets to one
0 0 1 0 A 0 A 0 z Copies A
0 0 1 1 A 0 A+1 c z Increments A
0 1 0 0 0 B B 0 z Copies B
0 1 0 1 0 B B+1 c z Increments B
0 1 1 0 A B A+B c z Adds A and B
0 1 1 1 A B A+B+ c z Adds A and B
1 and increments
1 0 0 0 0 B B 0 z One’s
complement of B
1 0 0 1 0 B B+1 c z Two’s
complement of B
(continued)
292 14 Calculators and Computers

Table 14.2 (continued)


Command input bits Full adder Outputs Description
inputs
C3 C2 C1 C0 A b S Carry Zero
1 0 1 0 A B A+B c z A plus one’s
complement of B
1 0 1 1 A B A−B c z Subtract B from
A
1 1 0 0 0 1 −1 0 0 All bits set to 1
1 1 0 1 0 1 0 1 1 Sets carry and
zero
1 1 1 0 A 1 A−1 c z Decrements A
1 1 1 1 A 1 A 1 z Copies A with
carry set

Table 14.3 A summary of the available commands for the ALU of Fig. 14.3
Command Command mnemonic Description
C3 C2 C1 C0 Hex
0 0 0 0 00 CLRA Clear accumulator to zero
0 0 0 1 01 ONEA Set accumulator to 1
0 0 1 0 02 CLRC Clear carry (set carry to 0)
0 0 1 1 03 INCA Increment accumulator
0 1 0 0 04 MOVB Move B to accumulator
0 1 0 1 05 INCB Move B + 1 to accumulator
0 1 1 0 06 ADDB Add B to accumulator
0 1 1 1 07 ADPB Add B + 1 to accumulator
1 0 0 0 08 CMPB Move B to accumulator
1 0 0 1 09 NEGB Move −B to accumulator
1 0 1 0 0A ADCB Add B to accumulator
1 0 1 1 0B SUBB Subtract B from accumulator
1 1 0 0 0C SETA Set all bits of accumulator to 1
1 1 0 1 0D CLCA Clear accumulator, carry Set
1 1 1 0 0E DECA Decrement accumulator
1 1 1 1 0F SETC Set carry to 1

In addition to the command for the ALU, a microprocessor will have commands
that move values between other various locations (e.g., from a memory location to
the register B). Each command is specified using 0’s and 1’s. The 0’s and 1’s within
the processor are referred to as the machine code. The command mnemonics are used
14.4 Base-2 Multiplication 293

for human consumption and are used as part of assembly language programming.2
Assembly language has a direct connection to the 0’s and 1’s, but is somewhat more
readable. Most programming is done using a higher-level language, such as C/C++ ,
Python, etc., which serves as a more convenient, and consistent, interface to the user.
That higher-level language will ultimately be converted to the 0’s and 1’s appropriate
for the processor being used.

14.4 Base-2 Multiplication

Base-2 binary multiplication can be accomplished using the same basic procedure as
is used for base ten, however the multiplication table is greatly simplified to include
only 0 and 1.
To perform multiplication using digital logic, it is possible to make a single circuit
as long as the number of bits is not too large, however it is usually more expedient
to perform the operation serially—i.e., as a series of steps rather than all at once.
Consider the following description of multiplication that is more easily translated
into circuitry:
1. Set A and B to the numbers to be multiplied
2. Initialize C to be zero (“Reset” or “Clear” C)
3. Examine the LSB of B, If LSB is a 1, then add A to C
4. Now shift A left by one bit, B right by one bit (the previous LSB of B is lost).
5. If B is not zero, loop back to step 3 and do it again. Otherwise, go on.
6. C now has the result.

Example Multiply 19 by 9.

19 Base 10 = 10,011 Base 2; 9 Base 10 = 1001 Base 2.

Step A B C
1,2 10011 1001 0
3 10011 1001 10011 LSB of B is 1 → Add A
4,5 100110 100 10011 B /= 0
3 100110 100 10011 LSB of B = 0
4,5 1001100 10 10011 B /= 0
3 1001100 10 10011 LSB of B = 0
4,5 10011000 1 10011 B /= 0
3 10011000 1 10101011 LSB of B = 1 → Add A
4,5,6 100110000 0 10101011 B = 0, C has result

2The mnemonics here are for this simplified ALU for example only, and may not appear in any
established programming language.
294 14 Calculators and Computers

Fig. 14.4 An iterative multiplying circuit. The adder is initially cleared and the results of each step
are added to the previous result

Figure 14.4 is a partial schematic for a circuit that does this for four bits. Note
that four bits are shown for A, B, and for the Result. It is more typical to allow for
twice as many bits in the result as are used for the inputs—the product of two four-bit
values may require up to eight bits for the result.
Earlier computers implemented the multiply function using software rather than
hardware. Some modern microcontrollers sometimes also rely in whole or in part on
software multipliers.
So-called “floating point” real number calculations are done using base-2 scientific
notation and integer operations. The “point” referred to is the decimal point (which
is commonly still referred to as a “decimal point” even though the number is not in
base ten). For floating point values in base two the number is written as a mantissa
(a value that is between 0.5 and 1.0) and an exponent for the power of two. The
computer memory stores the mantissa and the power together in memory. Floating
point values typically use a more traditional minus sign (here a 0 or 1 to indicate
positive or negative) rather than the two’s complement method.

Example Write the square root of two using base-2 scientific notation.

To do this, subtract the largest power of two that gives a positive (or zero) remainder
and put a 1 in the appropriate place. Then decrease the power to the next largest
power, etc. Recall that
√ digits to the right of the decimal place correspond to negative
powers. In base 10, 2 = 1.414213562.
14.5 Some Recursive Computations 295

Largest power of 2 less 1.414213562 − 1 = 0.414213562 First digit is in the “20 ’s”
than 1.414… is 20 place = 1
Next largest power of 2 The next digit is 0
is 2−1 = 0.5, which is 1.0
too big
2−2 = 0.25 is less 0.414213562 − 0.25 = 0.164213562 The next digit is 1
1.01
2−3 = 0.125 is less 0.164213562 − 0.125 = 0.039213562 The next digit is 1
1.011
2−4 = 0.0625 is too big The next digit is 0
1.0110
2−5 = 0.03125 is less 0.039213562 − 0.03125 = 0.007963562 The next digit is 1
1.01101
2−6 = 0.015625 is too The next digit is 0
big 1.011010
2−7 = 0.0078125 is 0.007963562 − 0.0078125 = 0.000151062 The next digit is 1
less 1.0110101
2−8 = 0.00390625 is The next digit is 0
too big 1.01101010
etc.… 1.01101010…


Hence, 2 = 1.01101010 . . . in base 2. To put this into base-2 scientific notation
with a mantissa between 0.5 and 1 the value must be divided by 2, that is it is shifted
right by 1 digit, so the mantissa becomes 0.101101010… and the exponent, the
number of multiples of 2 needed to√restore the original decimal point, is 1. Hence,
in base-2 scientific notation, write 2 = 0.101101010 × 21 .
To add two floating point values, the decimal points need to be lined up first. To
multiply two floating point values, multiply the mantissas as if they were integers,
put the decimal point to the far left of the result, then add the exponents as integers.
When stored in computer memory, the first 1 to the right of the decimal point
does not need to be stored since it will always be present. If this shortcut is used to
gain one extra binary digit, a special code needs to be used to refer to the value zero.
Often another special code is used to represent “not a number” or NAN. Such a code
is generated, for example, if the computer is asked to divide a value by zero.

14.5 Some Recursive Computations

Wiring a single digital circuit to perform some operations is not cost effective in
many cases. Computers (and calculators) will often use recursive numerical compu-
tations to provide what is needed. These calculations can be implemented using a
combination of software and hardware.
296 14 Calculators and Computers

The example algorithms below are for computation using digital logic capable of at
least NOT, addition, and shift operations, from which one can also obtain subtraction
and multiplication. The methods shown here illustrate how these operations might be
accomplished and may not be the most effective to use in practice. Modern computers
may include built-in specialized circuitry to perform many of the more common
functions. That these methods work can be verified using a hand-held calculator.
There is certainly no need to memorize the details of these examples.

14.5.1 Compute 1/x

Division is often accomplished by multiplying by the inverse. To compute the inverse


of a value x, do the following:
(a) Write x in base-2 scientific notation: x = d × 2n , where 0.5 ≤ d < 1.
(b) Make a reasonable initial guess for 1/d, called c. For example, c = 1/0.75 =
1.333. (Or one can use a look-up table and use a more accurate guess depending
on the value of d).
(c) Set b0 = c, a0 = b0 × d.
(d) Compute

bi+1 = bi × (2 − ai ) and
ai+1 = ai × (2 − ai )

repeatedly (i.e., i = 0, 1, 2, 3, …) until the result, bi , does not change within


some specified accuracy (ai will converge to 1.000).
(e) Then lim bi = 1/d and so 1/x = (1/d) × 2−n .
i→∞

14.5.2 Compute (1/x)½

One way to compute square roots, is to compute the inverse of the square root, then
use the inverse function above.
(a) Write x as x = d × 2n , where 4 ≤ d < 16 and n is an even integer.
(This particular range for d is only necessary because of the specific initial guess
shown below). √
(b) Make a reasonable initial guess for 1/ d, called c. For example, c = 2.5/d.
(c) Set b0 = c, a0 = b × b × d/2
(d) Compute

bi+1 = bi × (1.5 − ai )
ai+1 = ai × (1.5 − ai )2
14.5 Some Recursive Computations 297

repeatedly until the result does not change


( √ within
) some specified accuracy.
√ √ −n/2
(e) Then lim bi = 1/ d and 1/ x = 1/ d × 2 .
i→∞

14.5.3 Compute x½

Method I:

(a) Compute inverse of 1/ x using the above routines, or
(b) Make an initial guess, for example, b0 = 1 or b0 = x. (This guess does not need
to be particularly good).
(c) Then compute

bi+1 = (bi + x/bi )/2

repeatedly until the changes in the result become smaller than some specified
accuracy. Then bi has the result. This latter method takes a relatively long time
since the inverse is used many times.3
Method II:
This is basically a search method that works in base-2 binary with a fixed number of
bits. If the number of bits is not too large, this method can be surprisingly fast since
no inverse operations are required. This uses a search method starting with the MSB.
(a) Start with y = 0.
(b) Change the most significant bit of y to 1. Compute y2 .
(c) If y2 > x, then change the bit back to a 0, otherwise leave it a 1.
(d) Change the next most significant bit to a 1. Compute y2 .
(e) Loop to step c until all the bits of y have been tried. The result will be y.

3Some interesting integer approximations can be obtained by following this method using pencil

and paper. For example, start with b0 = 1 and iterate by√hand to get 2 ≈ 577/408 accurate to about
6 decimal places. Do just one more iteration to show 2 ≈ 665,857/470,832 which is accurate to
about 12 decimal places.
298 14 Calculators and Computers

14.5.4 Compute tan(x)

The value of x should be expressed in radians and in the first octant (0 to 45°). (All
other angles can be put in this range and the result converted using trig identities.
Values of x outside this range may work, but possibly not as efficiently.)
(a) Store some initial values
(i) set a variable tiny to a very small non-zero number, say tiny = 1 × 10−30 .
(ii) store the value a = −x 2 for later use.
(b) set f 0 = x, c0 = x/tiny, d 0 = 1, b0 = 3.
(c) compute as follows, repeatedly until the result, f i , doesn’t change significantly
(or equivalently, the product ci d i is “close enough” to 1.000).
d i+1 = 1/(bi + a × d i ), if the denominator is zero, use d i+1 = 1/tiny
ci+1 = bi + a/ci , and if the result is zero, use ci+1 = tiny
f i+1 = f i × (ci+1 × d i+1 ), and if the result is zero, use f i+1 = tiny
bi+1 = bi + 2
Note: You can compute tan(x/2) and use trigonometric relationships to get tan(x),
sin(x), etc. for the entire first quadrant. Using a = + x 2 instead of −x 2 , yields the
hyperbolic tangent of x.
There are similar methods available to compute the inverse tangent, and hence
all inverse trig functions, as well as for exponentials (ex ). For more information see
Press [2].

14.5.5 Compute K(k)

(The complete elliptic integral of the first kind).


Set x 0 = 1 − k, y0 = 1 + k and then iterate to compute the so-called “arithmetic
geometric mean,”

xn+1 = (xn + yn )/2


yn+1 = xn yn .

The difference between x and y becomes very small after surprisingly few iter-
ations. Then K(k) = π/(2x n ). The complete elliptic integral of the first kind shows
up in many physics problems. For example, when computing the period, T, for a
pendulum of length l when the amplitude, θ 0 , is not small, T = (2/π ) K(sin(θ 0 /
2))T 0 , where T 0 is the period for very small angles.
14.7 Tri-state Outputs 299

14.6 Communications

In addition to numerical computations, a computer is a device that sends data from one
place to another. Some of the locations are memory, some are arithmetic logic units
or displays, etc. That is, some locations will leave the data untouched while others
will act on it in some way. Thus, data communication is a major factor governing
how fast a computer can go.
Serial communications—data is spread out in time. Sends 1 bit at a time (includes
schemes for indicating “start” and “stop” of data). Examples: Serial (com) port on
computer, USB, most network communications including digital cell phones. As
digital electronics has gotten faster and faster, serial communications has become
quite common even between devices mounted on the same circuit board.
Parallel communications—data is sent simultaneously over parallel wires (or
parallel optical fibers, etc.). Send n bits at a time (plus control signals). Examples:
Printer port, IEEE-488. Parallel communication can be much faster, but as the number
of bits grows, the hardware must grow along with it and can become cumbersome.
Mixed—a mixture of parallel and serial. Example: dual tone numbers used for
phone dialing.

14.7 Tri-state Outputs

In order to facilitate the rapid movement of data from one place to another, a third
possible output is added to the digital logic circuits that is a “high impedance state,”
often labeled “Z.” Such devices are referred to as being “tri-state” or “3-state.”
Basically, the output is simply turned off, or effectively disconnected, in this third
state. In that way, millions of outputs can be wired together and as long as only one
of them is on at any given time, there will be no conflicts. The group of common
connections between the (potentially) millions of outputs is referred to as a “bus.”
A simple example of the use of tri-state outputs is shown in Fig. 14.5. Each circuit
connected to the bus has an “output enable” input. When that input is false, the device
cannot send data out to other devices—the output is disconnected. In Fig. 14.5, it is
assumed that when the output is disabled, the connections act as inputs. For some
devices, a separate “input enable,” or equivalent, may also need to be supplied.
Here, the “command” 011,000 (reading left to right) corresponds to “copy the
contents of A to B.” Similarly, 100,001 means “copy the contents of C to A.” Since
A, B, and C could be simple memory locations, inputs or outputs to an arithmetic
logic unit, a display device, etc., the ability to move data around using commands is
very powerful. Decoders and encoders (discussed in Chap. 12) can be used to make
the “command” more efficient and useful.
300 14 Calculators and Computers

Fig. 14.5 With tri-state


outputs, many circuits can be
connected to a common bus.
The output enable (oe) input
along with the clock input,
can be used to transfer data
from one register (set of
flip-flops) to another

14.8 Simplified CPU

The basic features of a computer’s central processing unit (CPU) are illustrated in
the diagram of a very simple (and hypothetical) CPU shown in Fig. 14.6. There are
a number of registers, to store intermediate values, an ALU to do some numerical
work, and control circuitry to run it all. Data and commands are retrieved (i.e., moved)
from external memory locations and data may be sent to memory, other processors,
displays, etc.
Program execution is accomplished by the CPU by, quite literally, simply moving
data from one place to another in a prescribed sequence. Some of the locations are
simply for storage, others cause some action to take place.

14.9 Other Uses for Tri-state Devices

The output of a typical tri-state device will be 0 V, a fixed voltage, V0 (such as 3.3 V
or 5 V), or a very high-impedance state, “Z”. A very high impedance is equivalent
to an open circuit. While perhaps not originally developed for such uses, that third
output can be used in some clever ways.

14.9.1 Measuring a Small Capacitance

A simple circuit that can be used to measure a small capacitance is shown in Fig. 14.7,
where C x is the small valued capacitor to be measured and C 1 is a much larger
capacitor of known value.4 To make a measurement, first S 2 and S 3 are closed to
discharge both capacitors. Then, with S 3 left open, switches S 1 and S 2 are alternately
closed and opened, taking care not to have them both closed at the same time.
After a number of cycles, the voltage across C 1 is measured with switch S 2 closed.
Alternatively, the number of cycles for the voltage across C 1 to reach some preset

4 See, for example, Philipp [1], where this technique is used as part of a proximity detector.
14.9 Other Uses for Tri-state Devices 301

Fig. 14.6 A simplified CPU

Fig. 14.7 A simple


switching circuit that can be
used to measure a small
capacitance

value, as determined by a comparator, can be used as a measure of C x . The basic


operation is the same as determining the size of a small bucket by measuring how
many times the small bucket must be filled in order to fill a larger bucket. This method
is related to the switched capacitor circuit mentioned in Chap. 4.
The capacitors are in series so the effective capacitance when S 2 is open is C eff =
C 1 ||C x ≈ C x . Now if C 1 is initially charged at a potential V i with charge Qi = C 1 V i
302 14 Calculators and Computers

and C x has been previously discharged through S 2 , then when S 2 is reopened and S 1
closed, the extra potential difference applied across the series capacitors is V 0 − V i ,
placing an additional charge of Qi+1 = C eff (V 0 − V i ) on C 1 . If this occurs over a
time /\t, then the average current into C 1 is I = /\Q//\t = C eff (V 0 − V i )//\t which
looks like a simple RC circuit with R = /\t/C eff . So, starting with the capacitors
discharged, the number of switching cycles necessary to get the potential on C 1 up
to a preset threshold can be used to determine C x .
Since the switches are either open or are connected to either a fixed voltage or to
ground, such a circuit might be implemented using tri-state digital logic rather than
switches. A simple tri-state buffer is shown schematically as a buffer with an extra
control input (Fig. 14.8). The behavior is summarized in Table 14.4. The output “Z”
corresponds to the high impedance state and for this buffer, a n , would correspond to
“output enable.” When the output is enabled, the output is either a fixed voltage (e.g.,
3.3 V or 5 V) or ground (0 V), depending on the input bn . The switched capacitor
circuit can thus be implemented by replacing the switches with tri-state logic and the
control signals, an , are used to switch the circuits on and off. In Fig. 14.9, a “high”
on the control signal opens the switch and a “low” closes it. If necessary, a small
resistor can be placed in series with the logic output in order to keep the current from
becoming too large. Such capacitance measurements are easily implemented using
a programmed microcontroller.

Fig. 14.8 Schematic for a


tri-state buffer

Table 14.4 Truth table for


an bn on
tri-state buffer, Fig. 14.8
L L L
L H H
H X Z

Fig. 14.9 The circuit of a1 a3


Fig. 14.7 implemented using
tri-state buffers in place of
switches Logic C1 a2
High
Cx
14.9 Other Uses for Tri-state Devices 303

14.9.2 Charlieplexing

Another clever application for tri-state outputs is called “Charlieplexing.” Char-


lieplexing, named for its inventor, Charles Allen, is a way of illuminating LEDs one
at a time. It is an efficient method, in terms of resource usage, to light LEDs directly
from microcontroller tri-state outputs. With n outputs from the microcontroller, you
can individually address as many as n2 − n LEDs without the use of any external
decoder circuitry. For example, 8 outputs could individually illuminate up to 56
LEDs.
The basic idea relies on the tri-state high impedance output and on the fact that
LEDs have a significant turn-on voltage. Between every pair of outputs, there are
two LEDs wired in opposite directions. In use, one output will be H (1), one L (0),
and the rest will be high impedance (Z).
Figure 14.10a illustrates the use of 3 outputs to address 6 LEDs. In order to
illuminate LED D2 , for example, set O1 to high impedance, O2 high and O3 low
which gives the equivalent circuit (where NC means “No Connection’) shown in
Fig. 14.10b. Diodes D1 , D5 , and D6 are clearly in reverse bias and will not illuminate.
There is a forward path from O2 to O3 through diodes D4 and D3 . However, the voltage
across that pair is limited by the turn-on voltage of D2 , a single LED, and is thus
insufficient to turn-on both D4 and D3 to any significant degree. Hence, they remain
dark. For this use, all the LEDs should be of a similar type, so they have a similar
turn-on voltage.
There may also be a set of outputs that can illuminate two LEDs at a time, though
clearly not all possible pairs can be simultaneously illuminated. For example, D1
and D4 could never be illuminated at the same time. The illusion that multiple LEDs
are lit simultaneously is created by rapidly switching through the appropriate output
states, illuminating the LEDs one at a time in such quick succession that they appear
to the eye to be on simultaneously.

Fig. 14.10 a The connections of LEDs to a microcontroller’s output lines for Charlieplexing. To
illuminate diode D2 , the outputs are set as shown in b
304 14 Calculators and Computers

14.10 Problems

1. Starting with a0 = 1, b0 = 2−1/2 = 0.70710678…, and s0 = 0, iterate a few times


using the following computations,

an−1 + bn−1 / ( )
an = ; bn = an−1 · bn−1 ; sn = sn−1 + 2n+1 an2 − bn2 ,
2
and when an and bn are close, compute

4an2
z= .
1 − sn

The result should be recognizable.


2. For the six-LED Charlieplexing example discussed at the end of this chapter, fill
in the missing values in the following table. Here “0” is ground, “1” is a positive
voltage and “Z” refers to the high-impedance state. (There is more than one way
to have no LEDs lit.)

O1 O2 O3 LED illuminated
None
1 0 Z D1
Z 1 0 D2
D3
0 1 Z D4
1 D5
0 D6

3. Write π in base-2 scientific notation to at least 8 places to the right of the base-2
decimal point.
4. What is the two’s complement of 0110100?
5. Compute −13 × 9 using 8-bit base-2 binary numbers and the 8-bit 2’s comple-
ment method to represent −13. Show that the 8 lowest bits of the result are the
2’s complement of 13 × 9 = 117.

References

1. H. Philipp, Charge Transfer Capacitance Measurement Circuit, US Patent 6,466,036 (1999)


2. W.H. Press, S.A. Teukolsky, W.T. Vetterling, and B.P. Flannery, Numerical Recipes, 2nd edn.
(Cambridge University Press, 1992)
Chapter 15
Quantum Logic

Quantum computation and quantum information utilize some of the rather strange
properties of quantum physics and have the potential to drastically change how some
digital circuits are implemented. This chapter is to provide a very brief introduction
into what is involved. No prior knowledge of quantum physics is assumed, though
that also means the discussion can go only so far. Quantum physics and quantum
computations are usually expressed using the mathematics of linear algebra—in
particular the language of vectors and tensors. Some prior exposure to this linear
algebra is assumed.
The hardware necessary for quantum logic and quantum computation is certainly
in its infancy. At the time of this writing, it is probably quite accurate to say that there
are many more experts in the use of quantum computation than there are quantum
computers. Even the simplest meaningful quantum computation can involve room-
sized, extremely expensive devices that are difficult to maintain. This is reminiscent
of the state of the room-sized classical computers of the 1950s. Nowadays there are
certainly many more classical computers than there are experts, and perhaps the same
will occur for quantum devices over the next few decades. This brief introduction is
included here in anticipation of such an occurrence.
There are many books and articles, at all levels, that describe quantum computa-
tion and quantum information.1 The intent here is to provide a brief introduction in
the context of the classical digital electronics already discussed. The details of any
specific quantum computation are considered beyond the scope of this presentation.
Quantum logic has some peculiar “needs” compared to classical logic. One of
them is that the logic gates should be reversible. To start, classical (transistor-based)
reversible gates are discussed. These can be constructed using the digital gates
discussed in Chaps. 12–14. A new notation and a new type of schematic diagram
are introduced, followed by a description of how basic quantum physics works and
a generalization of the classical ideas to quantum systems.

1 For example, see Bernhardt [1], Nielsen and Chuang [2]; and/or Singh et al. [3].
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 305
B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1_15
306 15 Quantum Logic

15.1 Reversible Gates

For the digital gates seen thus far the output is determined by knowledge of a number
of inputs. A reversible gate is one where knowledge of the outputs is also sufficient
to know the inputs. A simple NAND, with two inputs and one output, is clearly not
reversible. Knowledge of the (one) output does not uniquely determine the (two)
inputs. On the other hand, a simple inverter (NOT) is reversible and is its own
inverse—two NOTs in series will result in no change. In general, it would be expected
that to be reversible, a gate must have (at least) as many output values as input
values—no information can be lost.
A simple (non-trivial) reversible gate with two inputs and two outputs is a “con-
trolled not” or CNOT. Label the two inputs a and b, with corresponding outputs a'
and b' (see Fig. 15.1). Treating a as the control bit, which is unchanged, then the
outputs are a' = a and b' = a ⊕ b. That is, if a = 0, b' = b (“b is unaltered”) and if
a = 1, b' = b (b is inverted). Noting that a ⊕ (a ⊕ b) = (a ⊕ a) ⊕ b = b, it is clear
that this gate is its own inverse.
Two important reversible gates with three inputs are known as the Fredkin and
Toffoli gates. The truth tables for these two gates are shown in Table 15.1. It is
straightforward to show that each of these gates is its own inverse.
Both gates have inputs that can be labeled a, b, and c, and three corresponding
outputs, a' , b' , and c' . For the Fredkin gate, c can be considered a control gate and is
left unchanged: c' = c. If c is 0 then a and b are left alone: a' = a, b' = b. However, if c
is 1, a and b are swapped: a' = b and b' = a. This gate is also referred to as a controlled
swap, or CSWAP. Note that only the information is swapped. Imagine two people

Fig. 15.1 A classical


controlled NOT

Table 15.1 Truth tables for the Fredkin and Toffoli gates
Inputs Fredkin outputs Toffoli outputs
a b c a' b' c' a' b' c'
0 0 0 0 0 0 0 0 0
1 0 0 1 0 0 1 0 0
0 1 0 0 1 0 0 1 0
1 1 0 1 1 0 1 1 1
0 0 1 0 0 1 0 0 1
1 0 1 0 1 1 1 0 1
0 1 1 1 0 1 0 1 1
1 1 1 1 1 1 1 1 0
15.2 Schematics 307

walking next to each other, each holding a card that is green on one side and red on the
other. Each has one side (red or green) facing up. At some time, they are told to glance
at each other’s card and simultaneously alter theirs, as necessary, to be the same as
what they saw. No physical exchange has occurred, only the information—which
side is up—changes.
For the Toffoli gate, a and b are both considered control bits and are left unchanged:
a' = a, b' = b. If either, or both, a and b are 0, c is left alone, c' = c. If both a and
b are 1, then c is flipped, c' = c. That is, for the Toffoli gate, c' = c ⊕ (a · b). The
Toffoli gate is sometimes referred to as a CCNOT, since it is essentially a controlled
NOT with two control signals.
A Fredkin gate with constant inputs a = 1 and b = 0 can be used as an inverter
(NOT) in that a ' = c. If b = 0, then b' = a · c, and if b = 1, a ' = a + c. Thus,
these gates can be used to create NAND and NOR operations, which is sufficient
so that any (classical) logic operation can be constructed from these gates. Such a
gate is considered “complete.” Likewise, both the (classical) Toffoli and CNOT gates
are also complete. The Fredkin gate has an additional property, convenient for some
applications, that the number of 1’s in the output is the same as for the input—the
number of 1’s (and hence 0’s) is conserved. Such a gate is said to be “conservative.”
With these reversible gates, there must be (at least) as many outputs as inputs even
though only a small number of the outputs might actually be of interest. The “extra”
outputs, necessary for reversibility but not of interest in the end, are often referred
to as “garbage” output bits.
For quantum gates, there can be no “fanout,” where one output splits and becomes
the simultaneous input for multiple gates that follow, and hence the number of outputs
will equal the number of inputs.

15.2 Schematics

An alternate graphical representation is used when describing quantum logic. The


schematics for the three gates discussed so far are shown in Fig. 15.2.2 These
schematics are more abstract than those generally used for electronics. The data
values, from beginning to end, are shown with horizontal lines. They are read from
left to right. A vertical connecting line is shown where there is an interaction involving
data values. The horizontal lines do not (usually) represent wires, but describe a
sequence of events that occur with time from left to right—they are time lines.
For quantum computing, the data is the state of some physical system (atoms,
photons, nuclei, etc.) and the interactions are often caused by, or at least assisted by,
an outside influence, such as a laser pulse, electric field, or other applied perturbation.
The net result depends on the physical system, the sequence of events, and the details
of externally applied perturbations. Since the schematics are not specific to any
particular physical setup, they can be used to discuss and describe quantum logic

2 There are alternate versions of these in use.


308 15 Quantum Logic

Fig. 15.2 Schematics for a the CNOT, b Toffoli (CCNOT), and c Fredkin (CSWAP) gates

even in the absence of any practical quantum device. Note that combinations of these
gates are sometimes referred to as circuits, in analogy to the transistor-based logic
in common use, even though there may be no actual circuit involved.
In these schematics, a simple dot indicates a signal that is involved in an inter-
action, but is unaltered by the interaction, often referred to as a control signal. The
XOR symbol (⊕) is suggestive of the XOR operation, such as shown in Fig. 15.1
for the CNOT. The cross, with lines at 45°, is suggestive of the swap operation.
Other symbols are used for other interactions, and not all have been standardized. If
a vertical line crosses a horizontal line, but no dot or other symbol is present, then
those two systems do not interact.
As an example, consider the “circuit” shown in Fig. 15.3. It consists of a CNOT
followed by a Toffoli gate (CCNOT), followed by another CNOT. A classical real-
ization using (transistor-based) gates and wires is also shown. It is left as an exercise
to show that the combination of three gates in Fig. 15.3a is equivalent to a Fredkin
gate (CSWAP). A Toffoli gate can be constructed using six CNOT operations.

15.3 Tensor Notation

Consider the truth table for the CNOT gate, shown in Table 15.2. There are four
possible input pairs and four possible output pairs. Since the operation is reversible,
it must be the case that no output state is repeated. If the same output state results
from two different input states, the result is not reversible since knowledge of the
output cannot be used to uniquely determine the input. Hence, the outputs can only
be a rearrangement of the four possible inputs. The CNOT operation can thus be
described using simple vectors and tensors. While not very efficient for classical
logic gates, this notation, somewhat generalized, will be helpful for understanding
some quantum operations.
Describing the four input states, {(a,b)}, as unit column vectors, in the order
shown in the truth table,
15.3 Tensor Notation 309

Fig. 15.3 An example circuit shown using two different types of schematics. The schematic at
(a) generalizes for use with quantum logic, whereas (b) shows a classical realization using standard
transistor-based digital circuits

Table 15.2 The CNOT truth


Inputs Outputs
table
a b a' b'
0 0 0 0
0 1 0 1
1 0 1 1
1 1 0 1

⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞
1 0 0 0
⎜0⎟ ⎜1⎟ ⎜0⎟ ⎜0⎟
⎜ ⎟, ⎜ ⎟, ⎜ ⎟, and ⎜ ⎟, (15.1)
⎝0⎠ ⎝0⎠ ⎝1⎠ ⎝0⎠
0 0 0 1

the CNOT operation can be specified using the tensor, C NOT ,


⎛ ⎞
1 0 0 0
⎜0 1 0 0⎟
CN OT =⎜
⎝0
⎟. (15.2)
0 0 1⎠
0 0 1 0

The column vectors serve as a “basis” for the description in much the same way
that spatial coordinate axes are used to describe positions, velocities, and the like in
mechanics.
310 15 Quantum Logic

For example, if the input state is a = 1, b = 0, which is the third entry in the table,
then the output is found from
⎛ ⎞⎛ ⎞ ⎛ ⎞
1 0 0 0 0 0
⎜0 1 0 0 ⎟⎜ 0 ⎟ ⎜ 0 ⎟
⎟⎜ ⎟ ⎜
⎜ = ⎟, (15.3)
⎝0 0 0 1 ⎠⎝ 1 ⎠ ⎝ 0 ⎠
0 0 1 0 0 1

which is the fourth entry, so, referring back to the (first column of the) table, the
outputs are a' = 1, b' = 1. The tensor “operates” on the input vector to produce the
output (column) vector. Knowledge of the tensor (the “operator”) is equivalent to
knowledge of the truth table.
Similarly, the results of the Fredkin and Toffoli gates can each be described using 8
× 8 tensors, and more generally, with N inputs, the total interaction can, in principle,
be described by a 2N × 2N tensor. In many cases, only a subset of the inputs take
part in any specific interaction, so a smaller tensor can be used for those, and the
remaining inputs are simply carried forward unchanged.
In addition to column vectors, row vectors will be important.3 “Multiplying” a row
vector and a column vector (in that order), sometimes called a vector dot product,
results in a scalar (i.e., a single number). For example,
⎛⎞
b1
( )⎜ b2 ⎟
a1 a2 a3 a4 ⎜ ⎟
⎝ b3 ⎠ = a1 b1 + a2 b2 + a3 b3 + a4 b4 . (15.4)
b4

Since the output is no longer a vector, the output cannot be used as an input for
a logic gate. Note also that with this representation, if the order on the left side is
changed, so that the column vector is first, the result is a tensor, a completely different
beast!
The operators used for quantum logic are unitary. That is, they preserve the result
of a dot product (i.e., “lengths”). A rotation of the entire system is one example of
an operation that does not change dot products. Another example is a reflection, or
mirror image, operator.
An alternate, more compact notation is routinely used to describe these vectors
for quantum physics. Each of the basis vectors is described using a “-ket,” and an
identifier is included inside the -ket. For the CNOT, used above as an example, the
four basis vectors might be written

|00>, |10>, |10>, and |11>,

3 More technically, the row vectors are “dual vectors” for the corresponding column vector. Highly
technical definitions and language from mathematics is avoided here in an attempt to keep the
presentation from getting bogged down by the language. In the process, some loss of rigor may
occur.
15.4 Quantum Physics 311

where the indices here are the inputs a and b. The corresponding row vectors become

<00|, <01|, <10|, and <11|,

and are each referred to as a “bra-,” so the combination is called “bra-ket” notation
or “Dirac notation.” Any general vector can then be described as a combination of
the unit vectors. For example,

|B> = b1 |00> + b2 |01> + b3 |10> + b4 |11>, (15.5)

where {bi } are numerical constants. The general case is where it gets interesting.
Going back to the original (classical) digital logic problem, the inputs can be either
zero or one (i.e., false or true). For logic, what can such a general vector possibly
represent? It appears to represent situations that are some combination of 0 and
1, or both true and false at the same time. Such is one of the weird behaviors of
quantum mechanics, to be considered in more detail later, and so describing this
mixed situation is actually useful, and indeed necessary, for quantum logic.
The vector dot product is written as a bra- followed by a -ket. The vector dot
product between basis vectors gives 0 unless the indices are the same, in which case
the result is 1—they are an orthonormal basis set, like the directional unit vectors
from introductory physics. In the general case, shown with 4 basis vectors,

<A | B> = (a1 |00> + a2 |01> + a3 |10> + a4 |11>) (b1 |00> + b2 |01> + b3 |10> + b4 |11>)
= a1 b1 + a2 b2 + a3 b3 + a4 b4 . (15.6)

The process easily generalizes to any number of basis vectors. These dot products
also play an important role when describing quantum physics.
If an operator is present between the bra- and -ket, such as <A|C N O T |B>, the
operator can either operate on the vector to the left or on the vector on the right
(but not both) using the usual rules for tensors and vectors, and then the dot product
between the resulting states is computed.

15.4 Quantum Physics

The need for quantum physics arose near the end of the 1800s and beginning of the
1900s. Attempts to understand measurements of black body radiation using clas-
sical physics led to the so-called ultraviolet catastrophe—that the best explanation
required an unlimited amount of energy in the ultraviolet wavelengths. The idea that
light comes in discrete packets, now called photons, helped solve that problem. If
the (classical) wavelength of the light is λ, the photons each have an energy of hc/
λ—nothing more, nothing less—where h is Plank’s constant. Photons arose again
312 15 Quantum Logic

in Einstein’s explanation of the photoelectric effect.4 Thus, light came to be under-


stood as having both wave-like and particle-like properties, depending on the type
of measurement being made. During the same time, so-called cathode rays,5 created
electrically in the lab, were found to be a beam of electrons—the same lightweight
particles that were an important part of atoms. Not much later, those electrons were
found to also have wave-like properties.
The wave nature of electrons, and other particles with mass, is often described
using Schrödinger’s wave equation, valid only for non-relativistic particles.
While much of introductory quantum physics is devoted to specific solutions of
Schrödinger’s wave equation, the basic behavior necessary to understand the mathe-
matics of quantum physics that is required here can be described using photons and
simple measurements that can be performed using light.
Classically, light waves are described as traveling waves of electric and magnetic
fields, where both fields are perpendicular to the direction of travel and to each other.
By convention, the polarization of the light is determined by the direction of the
electric field. If the electric field oscillates in the vertical (horizontal) direction, the
light is said to be linearly polarized in the vertical (horizontal) direction. Light of
any polarization can be described as a sum of vertical and horizontal components.6
Light with vertical and horizontal polarizations form a basis that can be used for all
polarizations in much the same way that unit vectors in the x- and y-directions can
be used to describe all vectors in the x–y plane. Of course, those two perpendicular
directions (vertical and horizontal or x and y) can be rotated by any amount and any
polarization (or vector) can be described using those new directions—they form a new
basis. What is considered vertical and horizontal is determined by the experimenter,
not by the light.
A device that admits only one direction of polarization is called a polarizer. If light
is sent through an ideal vertical polarizer, any light that gets through will emerge with
vertical polarization. Similarly, light that emerges from any (ideal) polarizer oriented
in any specified direction will be light of only that polarization.7 What will be most
interesting is that if unpolarized light, such as light from an incandescent lightbulb,
is sent through a polarizer, the light that emerges has half the intensity of the incident
light.
A usual set of demonstrations is illustrated in Fig. 15.4. Light is clearly observed
to emerge from a first polarizer (Fig. 15.4a). A second polarizer can be used to show
that the light is polarized. When the second polarizer is aligned with the first, the light
makes it through. When the second polarizer is perpendicular to the first (Fig. 15.4b),

4 Einstein’s only Nobel prize was for his explanation of the photoelectric effect, despite his many
other major accomplishments.
5 In addition to cathode rays, other measurements at the time resulted in what were described as

beta rays, delta rays, epsilon rays, and Lenard rays, all of which turned out to be beams of electrons.
6 Circular polarization is excluded from the discussion here as being unnecessary for the immediate

topic at hand.
7 Henceforth, only ideal polarizers are considered.
15.4 Quantum Physics 313

Fig. 15.4 Experimental


arrangements involving (a)
a one, b two, and c three
polarizers

(b)

(c)

the light is blocked. As the relative angle of the polarizers changes from 0° to 90°,
the light that emerges gradually goes from a maximum to zero.
A third polarizer between two perpendicular polarizers (Fig. 15.4c), when oriented
at an intermediate angle is found to allow some light to emerge from the final polarizer
even if it is perpendicular to the first. This seemingly magical result is easily explained
classically.
The classical description is that the light can be considered to be made up of a
combination of light polarized parallel and perpendicular to the polarizer direction,
and only the component parallel to the polarizer makes it through. For the unpolarized
light there is as much parallel as perpendicular so half the intensity makes it through.
If polarized light is incident on a second polarizer, the wave is rewritten in terms of
the parallel and perpendicular components for the second polarizer, a new basis, and
the parallel component (if any) makes it through. The process is repeated for any
subsequent polarizers.
Now, consider that the light comes in discrete particles of energy. In this process,
fractional photons, such as half-photons, are never observed. Either a photon is
observed or nothing.
If light comes in photons, and unpolarized light is sent through a vertical polarizer
where half the energy makes it through, it must be that half the photons make it
through—the energy per photon does not change. Once the photons emerge, they
must have the polarization of the polarizer. One is tempted to say that the light
incident on the polarizer must have had half the photons polarized vertically and half
horizontally. But now if the single polarizer is rotated, say by 45°, it is still the case
that half get through, and the half that get through are polarized at 45°. If initially
half the photons are vertically polarized, and half are horizontally polarized, how
can any make it through the rotated polarizer? Did the photons before the polarizer
change because the experimenter rotated the polarizer? It seems that it is not easy
to deduce the polarization of the photons before the polarizer from a measurement
made after the polarizer.
Consider now the setup of Fig. 15.4c where unpolarized light is incident on a
vertical polarizer, then that light is incident on a second polarizer rotated 45° from
314 15 Quantum Logic

vertical. This is easy to do and to analyze using the classical wave description of light.
Half of the vertically polarized light intensity makes it through the rotated polarizer.
That is, half of the vertically polarized photons make it through the rotated polarizer
and all that emerge now have a polarization rotated 45° from vertical. If now a third
polarizer is added, rotated another 45° so it is now horizontal, half of the photons from
the middle polarizer will then make it through. Hence, there are some photons that
emerge with horizontal polarization despite the fact that all the horizontal polarization
was removed by the first polarizer. The very nature of the observed particles seems
to have been changed by the measurements of the experimenter.
So, how can the behavior of the discrete particles of light, the photons, be described
so that the result matches the observations? Classically, the light is written as a
combination of a component parallel and a component perpendicular to the polarizer
direction. A similar approach can be used to describe the photons.
If the propagation direction is z and the polarizer is directed along y, then the
perpendicular direction is x. A photon is then represented as a combination of basis
states that separately describe a photon polarized along x and y respectively. That is,
using |ψ> to represent the state of the photon, write8

|ψ> = a|x> + b|y>. (15.7)

Classically, the energy of an electromagnetic wave (e.g., light) is found from the
squares of the electric and magnetic fields. If a photon is observed, its energy is
known and fixed (and if it is not observed, the observed energy is zero). Looking at
the dot product of the photon with itself, which is, in some sense, its square,

<ψ|ψ> = a 2 <x|x> + b2 <y|y> = a 2 + b2 . (15.8)

and if this is considered the projection of the photon onto itself, then it must be that
a 2 + b2 = 1. This is referred to as the normalization condition. You get one photon
when you measure along that photon’s polarization direction. Never a fraction, never
more than one.
Now, consider the photon to be incident on a polarizer aligned along the y-
direction. The x-component will not make it through. An operator that can be used
to describe the process is constructed from the basis vectors. That is

<ψ| y><y| ψ> = b2 . (15.9)

Now, b2 ≤ 1, but you either see the photon or you do not. This leads to the
probabilistic interpretation that b2 is the probability that the photon is observed.
If the measurement is repeated many times, the fraction of the photons that make
it through matches the fraction of the energy that makes it through as computed
classically.

8In general, a and b can be complex values in which case the dual (row) vector uses the complex
conjugates a* and b*.
15.5 The Qubit 315

The examples above used, as examples, polarizers that block some of the photons.
Such a device would be undesirable for quantum computation due to the associated
loss. However, the polarizations can be separated in other ways, such as with a
birefringent material,9 where (ideally) no loss occurs. That is, an incident beam
can be separated into two beams, one of which is polarized vertically and the other
horizontally, without throwing away any photons.

15.5 The Qubit

The quantum bit, or qubit, is then a bit that is not just 0 or 1, but may be simultaneously
both. An association between some physical property, such as the polarization of
light, is used to represent 0 and 1. Any two-state system can be used. For example,
electrons have a spin that can be “up” or “down.” Up and down are again defined by the
experimental set-up, usually by an applied magnetic field. Many atomic nuclei (i.e.,
isotopes) are also spin ½ and can be used, such as 1 H, 13 C, 19 F, and 29 Si. Many nuclei
have more than two states and in principle, systems with more than two states can be
used—in essence to have base 3, 4, 5, etc., rather than base 2 (binary) logic—though
the complications of doing so have not yet outweighed the extra difficulty of using
systems with more than two states. Superconductivity is a quantum phenomenon and
arrays of superconductive materials show promise for some large-scale use. Various
other schemes include the use of interactions between electrons and trapped ions.
Many of the current schemes need to be operated near absolute zero to work.
No matter what two-state system is used, once the association between the two
physical states, |0> and |1>, and the labels 0 and 1 has been made, the quantum bit,
or “qubit,” can be described as

|ψ> = a|0> + b|1> (15.10)

where a and b are (complex) constants and |a|2 + |b|2 = 1. Using vector notation,
the qubit would be written
( ) ( ) ( )
1 0 a
a +b = . (15.11)
0 1 b

Each qubit in a system represents the current state of a particular (two-state)


physical system. A measurement of the system will yield |0> a fraction |a|2 of the
time, and |1> a fraction |b|2 of the time. Once measured, the system stays in the
observed state.

9 Calcite crystals are commonly used to demonstrate this effect.


316 15 Quantum Logic

(a )
Table 15.3 Results of single qubit operations on |ψ> = a|0> + b|1> or b

Operation Result
Name Tensor representation Vector representation Dirac notation
( ) ( )
I or identity 10 a a|0> + b|1>
01 b
( ) ( )
X or NOT 01 b b|0> + a|1>
10 a
( ) ( )
Z 1 0 a a|0> − b|1>
0 −1 −b
( ) ( ) √
H or Hadamard 1 1 a+b ((a + b)|0> + (a − b)|1>)/ 2
√1 √1
2 1 −1 2 a−b

There are several operations involving a single qubit that are useful. A few of these
operations are summarized in Table 15.3. In a schematic, these are often signified
using a box containing an identifying letter.
The Hadamard gate, related to the 45° rotation in the polarization discussed above,
is useful for creating entangled states as will be illustrated in the next section. The
NOT and X operator are the same. Which name appears depends on circumstances.10
As was seen for the CNOT, all of these single qubit operators can (in principle) be
controlled by one or more other qubits.
In general, a single qubit operator must be unitary. A general unitary operator can
be written
( )
cos(θ/2) −eiβ sin(θ/2)
U (θ, α, β) = . (15.12)
eiα sin(θ/2) ei (α+β) cos(θ/2)

The inverse of U is the transpose of U (the rows and columns of U are switched).
The use of θ/2 in the trig functions, rather than θ , comes from the mathematics for
rotations of spin-½ particles. Such a particle is characterized as spin up and spin
down. Thus, a rotation of 180° changes from one state to the other—e.g., up to
down. Down, however, is not the same as up with a minus sign. Compare this to the
examples above using photons where a physical rotation of 90° changes from one
polarization to the other.11 This has an interesting effect that, for a spin-½ particle,
to get back to the starting point requires a 720° physical rotation. A 360° rotation

10 This arises from consideration of the so-called Pauli matrices. There is also a Y operator, similar
to X, except the non-zero entries are i and −i in place of the 1’s.
11 Photons are spin-1 particles. Spin-1 particles normally have three sub-states: −1, 0, and +1. Due

to relativistic effects, the state 0 is not observed for particles traveling at the speed of light. Hence
photons look like a two-state system.
15.6 Entanglement 317

changes the state by a minus sign. Since only the square is observed, that minus sign
will not be directly observable.
Classical logic has a very limited number of possible inputs and outputs compared
to the superpositions allowed by quantum mechanics. Some quantum computations
will use this to solve problems using a basis change (e.g., a rotation or other unitary
transformation) before an interaction. This is, crudely speaking, somewhat analogous
to the use of rotated x–y coordinates used to conveniently solve mechanics problems
involving an object sliding or rolling down an incline.

15.6 Entanglement

In general, the state, |ψ>, for a two-qubit system (labeled a and b) can be written as

|ψ> = (A|0>a + B|1>a ) (C|0>b + D|1>b ), (15.13)

where the (possibly complex) constants satisfy

|A|2 + |B|2 = 1; |C|2 + |D|2 = 1. (15.14)

As an example, consider the case where



A = B = 1/ 2; C = 1; D = 0. (15.15)

Then the combined system is described by


( ) ( )
1 1
|ψ> = √ (|0>a + |1>a )|0>b = √ (|0>a |0>b + |1>a |0>b ). (15.16)
2 2

There is equal probability that the states |0>a |0>b and |1>a |0>b will be observed.
That is, qubit b is always observed as |0>b and there are equal probabilities that the
control qubit, a, will be observed in either state. Knowledge of the results for either
a or b tells you nothing about the other. That should not be too surprising.
Now send that state through a CNOT quantum gate, in this case with a used as
the control qubit, to get
( )
| '> 1
|ψ = √ (|0>a |0>b + |1>a |1>b ). (15.17)
2

Now if a is observed to be in the state |0>a , then b will be observed in state |0>b
and if a is observed to be in the state |1>a , then b will be observed in state |1>b and
vice versa. Measurement of one determines the other. In situations like this example,
where the results of measurements of separate systems become correlated, the states
are said to be “entangled.” What is most interesting is that, if care is taken to avoid
318 15 Quantum Logic

|0 M
H
|0

Fig. 15.5 Starting with two prepared states, a Hadamard gate followed by a CNOT results in an
entangled pair so that measurement of one output determines the other

outside influences on the system, the two measurements can be well separated in both
time and space. Whoever measures first will instantaneously determine the result for
the other.
The process can be implemented as shown in Fig. 15.5. The two inputs are initially
prepared, in this case both in the 0 state. One of the inputs (a) is sent through the
Hadamard gate to create the superposition that is used as the control signal. The two
outputs will then be entangled. The device shown at the end, intended to represent
a meter of some kind, symbolizes a measurement of the qubit. Once the qubit is
measured it will become either a 0 or 1 and the superposition is lost. You cannot
tell what the state was before the measurement. The double line following the meter
indicates the data is now a classical bit (either 0 or 1). In this case, the measurement
(shown for a ' ), which has an even chance of being 0 or 1, will also determine the
state of b' .
Note that many authors will simplify the notation in a number of ways so that

|ψ> = (A|0>a + B|1>a ) (C|0>b + D|1>b )


= AC|0>|0> + AD|0>|1> + BC|1>|0> + B D|1>|1>
= AC|00> + AD|01> + BC|10> + B D|11>
= r |00> + s|01> + t|10> + u|11>. (15.18)

With this notation, the order of the indices must be respected—in this case the
first is for qubit a, the second for b. The state |0>|1> is not the same as the state |1>|0>.
The possibility of having entangled quantum states, a non-classical result, gives
rise to one major difference between some of the capabilities of quantum and classical
logic, especially for information transmission.

15.7 The Nature and Future of Quantum Computations

Quantum computations are known to have great advantage for some specific prob-
lems. One of particular interest is to determine the prime factors of large numbers. It
is relatively easy to multiply two prime numbers to get a composite number. For such
a problem, there is likely little, if any, advantage to use a quantum method over clas-
sical methods. It is much harder—many more steps are involved—to find the prime
factorization of a composite number. This is especially true as the number of digits
15.8 Problems 319

grows. The scaling of (all known) classical computations to find the prime factors of
large integers is such that after a certain point, even the fastest (classical) computers
are no match for the problem. The scaling for quantum methods is different, and does
not, in principle, run into that limitation nearly as quickly. Here there is an advantage
for quantum methods. Factorization is an important problem because the security of
digital data depends on the fact that it is (currently) a tough problem. Other potential
uses include sorting algorithms (important for searches of large databases).
To take advantage of the scaling for large numbers (or large datasets) requires a
large number of interconnected qubits. One problem is that if all qubits are to be
connectable to each other, e.g., for use as control qubits, then there must be a huge
number of interconnections. Since interactions between qubits are often based on
close proximity (and not wires) construction of a large quantum processor becomes
problematic. Instead, qubits can be connected in subsets. Then the control over
which qubits are used for any particular problem requires careful management of
the resources. It may also be the case that there are problems that remain tractable
on classical computers that do not scale well for real quantum computations.
A most likely scenario is a hybrid device that uses classical devices (e.g., transis-
tors) to do much of the control and management with a quantum co-processor that
can be invoked only as necessary. Hybrid devices may also be necessary for some
digital circuits involving feedback, such as flip-flops, and for large-scale memory
storage.
At present, the problem of making a quantum computer and the problem of how
to get it to produce useful results are largely being researched separately. There is a
long way to go before the use of quantum devices will be routine.

15.8 Problems

1. Show that the outputs of the Fredkin gate (see Table 15.1) can be written

a ' = c · a + c · b; b' = c · a + c · b; c' = c.

2. Consider the Toffoli gate (see Table 15.1) with one or more inputs fixed to show
that it produces c' = a · b (AND) when c = 0, c' = a · b (NAND) when c = 1,
c' = a ⊕ c (XOR) when b = 1, and c' = a (NOT) when b = c = 1.
3. Fill in the steps to show that the circuits in Fig. 15.3 are equivalent to a Fredkin
(CSWAP) gate.
4. Show that
⎛ ⎛ ⎞⎞
b1
(( ) )⎜ ⎜ b2 ⎟⎟
a1 a2 a3 a4 C N O T ⎜ ⎜ ⎟⎟
⎝C N O T ⎝ b3 ⎠⎠ = a1 b1 + a2 b2 + a3 b3 + a4 b4 .
b4
320 15 Quantum Logic

(Hint: just consider what the C NOT does to each vector and (15.4). Do not spend
time writing out and multiplying tensors and vectors term by term.) This verifies
that C NOT is its own inverse.
5. What values of θ, α, and β should be used in (15.8) to get the
(a) NOT operator, (b) the Z operator, and (c) the H operator?
6. Write out the classical truth table for the circuit in Fig. 15.P6 and show that it
functions as a full adder for input bits a, b, and cin , with outputs sum s and carry,
cout . Include enough intermediate values in your truth table to provide convincing
evidence for your result.

Fig. 15.P6 Problem 6

7. Using (15.14) show that the new constants in (15.18) will satisfy

|r |2 + |s|2 + |t|2 + |u|2 = 1.

8. Show that the gate described by the tensor


( )
1 1+i 1−i
V =
2 1−i 1+i

describes the square root of the NOT operator. That is, applying V twice in a row
is the same as the NOT operation. Such an operation does not exist for classical
computation.

References

1. C. Bernhardt, Quantum Computing for Everyone, (MIT Press, 2020).


2. M.A. Nielsen, I.L. Chuang, Quantum Computation and Quantum Information, 10th Anniversary
Edition, (Cambridge Univ. Press, 2016)
3. C. Singh, A. Levy, J. Levy, Preparing precollege students for the second quantum revolution
with core concepts in quantum information science. Phys. Teach. 60, 639–641 (2022)
Appendix

A review of units and some of the mathematics used for electronics follows.

A.1 SI Units

Many of the standard international (SI) units that are often encountered in electronics
are summarized in Table A.1.

A.2 Common Unit Prefixes

Prefixes are used with units to convey a power of ten multiplier. Many are defined
and are commonly used with Standard International (SI) units. Some of the more
common prefixes encountered in electronics are listed in Table A.2.

Table A.1 SI units often


SI units (MKSA)
encountered in electronics
1 amp = 1 A = 1 C/sec
1 V = 1 m2 kg s−3 A−1
1  = 1 V/amp = 1 m2 kg s−3 A−2
1 S = (1 )−1 = 1 mho = 1 amp/volt
1 F = 1 amp sec/volt = 1 s/ = 1 m−2 kg−1 s4 A2
1 H = 1 V s/amp = 1 ·s = 1 m2 kg s−2 A−2
1 Hz = 1 repeat/s = 1 cycle/s = 1 cps = 2π rad/s
1 W = 1 V A = 1 N m/s = 1 m2 kg s−3

© The Editor(s) (if applicable) and The Author(s), under exclusive license 321
to Springer Nature Switzerland AG 2023
B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1
322 Appendix

Table A.2 Some of the more


Symbol Prefix Multiplier
common unit prefixes and
their meanings p pico- 10−12
n nano- 10−9
μ micro- 10−6
m milli- 10−3
c centi- 10−2
d deci- 10−1
k kilo- 10+3
M mega- 10+6
G giga- 10+9
T tera- 10+12

A.3 Fourier Series

Near the turn of the nineteenth century, Joseph Fourier (1768–1830), developed many
important theorems, including several related to vibrations. Fourier showed that for
any periodic signal, call it S(t), there is a choice for constants A0 , A1 , A2 , etc., called
amplitudes, and ϕ 1 , ϕ 2 , etc., called phases, such that



S(t) = A0 + An sin(nω0 t + ϕn ), (A.1)
n=1

where the periodic function has a repeat frequency f 0 = 2πω0 . The series of terms
on the right would be referred to as the “Fourier series representation of S(t).” With
minor adjustments, trigonometric identities can be used to find equivalent series
representations using cosine functions, a combination of sine and cosine functions,
or any equivalent, instead of sine functions. Aside from the use of these trigonometric
identities, the series representation of S(t) is unique.
In some cases, mathematical equality in (A.1) is only achieved in the limit that an
infinite number of constants and an infinite number of sinusoidal (e.g., sine) functions
are used. Despite that, when representing any measured signal, the result of the sum
is finite. A finite number of terms will usually suffice to achieve any desired level of
accuracy for any electronic signals measured in the real world.
It should be obvious that if enough of the constants are known, S(t) can be
computed for any time, t, using the equation above. All that is required is a calculator
with a “Sin” function. What is less obvious is that if S(t) is known for at least one
repeat cycle, it is possible to determine each of the constants for the series. In prac-
tice, if N distinct values of S(t) are known, equally spaced in time over one repeat
cycle, that is sufficient data to compute N constants. To determine those N constants,
a Fourier transform (FT) is used.
Fourier’s theorem also shows that there is no difference between a periodic func-
tion produced directly, whatever its origin, from the result obtained by adding a large
Appendix 323

number of sine functions, with appropriately chosen amplitudes and phase constants,
created by separate signal generators. Hence any periodic signal can be considered
to be made up of, or composed from, a summation of a number of sinusoidal signals,
no matter how that signal was actually generated in the first place. Using the super-
position principle, when it applies, each of those sinusoidal signals can be considered
separately and the results added together—the sum of the solutions is the solution to
the sum. Analyzing a circuit one sine function at a time is sometimes referred to as
analysis “in the frequency domain.”
Four well-known mathematical examples are shown in Fig. A.1: the sine, square,
triangle, and sawtooth functions. The functions shown each have amplitude 1 and
repeat with frequency f 0 . The Fourier series representations of these functions can
be written:
• Sine function:

S(t) = sin(ω0 t) (A.2)

• Square function:


4  1
S(t) = sin(nω0 t)
π n odd n
 
4 1 1
= sin(ω0 t) + sin(3ω0 t) + sin(5ω0 t) + · · · (A.3)
π 3 5

• Triangle function:


8  (−1)(n−1)/2
S(t) = 2 sin(nω0 t)
π n odd n2
 
8 1 1
= 2 sin(ω0 t) − sin(3ω0 t) + sin(5ω0 t) + · · · (A.4)
π 9 25

• Sawtooth function:


2  (−1)n
S(t) = − sin(nω0 t)
π n=1 n
 
2 1 1
= sin(ω0 t) − sin(2ω0 t) + sin(3ω0 t) + · · · (A.5)
π 2 3

where ω0 = 2πf 0 , as before.


There are several cases where care must be exercised when using Fourier series
representations and analysis in the frequency domain. In particular where there are
324 Appendix

Fig. A.1 Sine a square b triangle c and sawtooth d functions shown as functions of time, t, with
t in units of 1/f 0

non-linearities so that superposition no longer applies. Also, for a signal that is


changing in time—for example it is turned on and/or off—the addition of non-
periodic time dependence can also make matters much more complicated.

A.3.1 How a Fourier Transform Works

There is considerable mathematical theory behind the Fourier transform that goes
beyond this discussion. The principle that makes the transform work can be
understood by considering the trigonometric identities

sin(ω1 t)sin(ω2 t) = 21 (cos((ω1 − ω2 )t) − cos((ω1 + ω2 )t)) (A.6a)

cos(ω1 t) cos(ω2 t) = 21 (cos((ω1 + ω2 )t) + cos((ω1 − ω2 )t)) (A.6b)

sin(ω1 t) cos(ω2 t) = 21 (sin((ω1 + ω2 )t) + sin((ω1 − ω2 )t)). (A.6c)

When such terms are averaged over time, the result will be zero unless ω1 = ω2 . If
ω1 = ω2 , then the first two (but not the third) will each average to one-half (because
cos(0) = 1).
Now start with any signal, S(t), and multiply it by a sine (or cosine) function that
has any frequency, ω. Mathematically, that is equivalent to multiplying the Fourier
series representation of S(t) by that same sine, even if you do not know what that
representation is. That is1 :

sin(ωt)S(t) = A0 sin(ωt) + A1 sin(ω0 t) sin(ωt)


+ A2 sin(2ω0 t) sin(ωt) + · · · . (A.7)

Now average that result over time. On the right it can be seen that the expression
will average to zero unless ω is equal to one of the frequencies ω0 , 2ω0 , 3ω0 , … and

1The phase factors have been left off for clarity. More generally they are included for the signal, in
one way or another, and multiplication by a cosine, averaged over time, will also be necessary.
Appendix 325

so on. That must be true on the left as well. Now sweep ω starting from zero, taking
a new time average for each new value of ω. When ω = ω0 , the result is proportional
to A1 . Increase ω some more and the whole expression averages to zero until ω =
2ω0 , at which point the result is proportional to A2 , and so on. The time-averaged
response as a function of the variable ω is the Fourier transform (aside from an overall
constant) and it yields the coefficients, An .
If a signal is not periodic, but finite in time, then it can be treated as if the signal
were periodic over a longer time. That is, if S(t) is known for 0 < t < T, then it can
be taken to repeat for T < t < 2 T, and so on, for the sake of the calculation. That
idea can be formally extended to the limit that T → ∞, and thus all time-dependent
signals, S(t), can be represented by a sum of sinusoids.
With digital sampling of a (finite length) signal, and the use of a computer, the
transform can be obtained relatively quickly. If the data is equally spaced in time and
the total number of data values is chosen appropriately, a common choice being 2N
where N is a positive integer, an algorithm known as a fast Fourier transform (FFT)
can be used. A FFT uses a particularly efficient way to organize and perform the
numerical calculation.

A.4 Complex Numbers—A Review

A.4.1 The Basics

A complex number, Z, can be written using two real numbers, A and B,

Z = A +iB (A.8)

where i is the square root of minus 1 (i2 = − 1). Some authors will use j instead of
i. A is called the “Real Part” and B the “Imaginary Part” of the number. The “Re”
and “Im” functions take a complex number as an argument and return the real and
imaginary parts respectively. That is

Re(Z ) = A; Im(Z ) = B. (A.9)

Using the identity2

ei x = cos(x) + i sin(x) (A.10)

it is possible to write

Z = Ceiϕ , (A.11)

2 Here “e” is Euler’s number, 2.7182818….


326 Appendix

Fig. A.2 The relationship


between real and imaginary
parts, A and B, and the
magnitude and phase, C and
ϕ, for a complex number
visualized using an
x–y graph

where C and ϕ are both real numbers. The magnitude, C, is related to the real and
imaginary parts through the Pythagorean theorem,

C = |Z | = A2 + B 2 , (A.12)

and

A = C cos(ϕ); B = C sin(ϕ), (A.13)

so

tan(ϕ) = B/A, or ϕ = tan−1 (B/A). (A.14)

Note that A and B look like the x and y components of a 2-dimensional vector
of length |Z|. Even though Z is not a vector, it is often convenient to think of it
as a vector with the real part corresponding to the x-component and the imaginary
part corresponding to the y-component. The phase angle, ϕ, would then be the angle
counter-clockwise from the x-axis. See Fig. A.2.
Multiplication and addition of complex values proceeds as for real numbers,
though with the inclusion of the definition of i. Hence if Z = A + iB and Y =
C + iD are two complex numbers, then

Z + Y = (A + C) + i (B + D)
(A.15)
Z · Y = (AC − B D) + i( AD + BC).

Note that multiplication is not the same as a dot product for vectors, though it
may be tempting to treat it so. Multiplication and division are often easier using the
magnitude and phase. For example, if

Z = |Z |eiϕ ; Y = |Y |ei θ , (A.16)

then the product is simply

Z · Y = |Z ||Y |ei(ϕ+θ ) . (A.17)


Appendix 327

Note that the amplitudes multiply and the phases add.


Multiplication and addition of complex values have the usual commutative,
distributive, and associative properties that are present for real numbers.
The complex conjugate of a complex number Z is designated with an asterisk,
Z * . Compute the complex conjugate by changing all “i’s” to “ −i’s” or change the
sign of the phase angle, ϕ. For example, if Z = A + iB = Ceiϕ then Z * = A − iB =
Ce−iϕ . One consequence of this definition is that

|Z |2 = Z · Z ∗ = A2 + B 2 . (A.18)

This last relation is often used as a convenient way to find |Z|.

A.4.2 Some Complex Identities

Some identities that are relatively easy to derive are:


 25
i 2 = −1, i 3 = −i, i 4 = 1, . . . , i 100 = i 4 = 1, i 101 = i, etc. (A.19)

ei π = −1, e2iπ = 1, etc. (A.20)

1/i = −i (A.21)

 ix   ix 
sin(x) = 1
2i
e − e−i x ; cos(x) = 1
2
e + e−i x (A.22)

e Z +Y = e Z · eY (A.23)

where Z and Y are any complex values and x is any real value.
Index

A Binary code
Active region non-weighted, 270
BJT, 156 weighted, 267
FET, 137 Binary Coded Decimal (BCD), 266
Alternating Current (AC), 52 Binary counter, 257
Ammeter, ideal, 8 Bipolar Junction Transistor (BJT), 155
Amplifier DC current gain, 157
automatic gain control (AGC), 230 model, 158
complementary symmetry, 181 saturation region, 172
differential, 181 BJT amplifiers
op-amp, 189 common base, 165
transistor, 145, 165 common collector, 165
Analog to digital conversion, 281 common emitter, 165
analog to Gray code, 283 Boolean algebra, 235
quantization noise, 284 basic results, 237
search scheme, 282 DeMorgan’s theorems, 237
timing schemes, 282 Boolean operations, 235
voltage to frequency, 281 Bra-ket notation, 311
Angular frequency, 53 Branch method, 13
Arithmetic Logic Unit (ALU), 290
ASCII, 271
Audio taper, 23 C
Cables. See transmission lines
Capacitors
B parallel and series, 64
Balanced code, 269 Cascode pair, 179
Bar codes, 271 Central Processing Unit (CPU), 300
Barkhausen criterion, 214 Charlieplexing, 303
Bias, forward and reverse, 118 Circuit
Biasing, 163 definition, 3
Biasing resistors, 163 open, 3
Bias line Circuit reduction, 9
BJT, 162 Code conversions
Binary, base-2, 265 base-2 binary to Gray code, 274
addition, 287 BCD to binary, 276
multiplication, 293 binary to BCD, 277
© The Editor(s) (if applicable) and The Author(s), under exclusive license 329
to Springer Nature Switzerland AG 2023
B. H. Suits, Electronics for Physicists, Undergraduate Lecture Notes in Physics,
https://doi.org/10.1007/978-3-031-36364-1
330 Index

decimal to Gray code, 275 Diode limiter, 123


Gray code to base-2 binary, 275 Dirac notation. See Bra-ket notation
Common Mode Rejection Ratio (CMRR), Direct Current (DC), 52
182 Dissipation angle, 75
Communications Dissipation factor, 75
parallel, 299 Duality, 95
serial, 299 Duty cycle, 227, 278
Comparator, 222 Dynamic memory, 151
with hysteresis, 223
Comparator circuits
level indicator, 224 E
oscillator, 226 Earth. See ground
pulse conditioner, 229 Effective resistance, 8
pulse generator, 225 parallel resistors, 8
zero crossing detector, 228 series resistors, 8
Complementary symmetry amplifier, 181 Electric eye, 175, 254
Complex number, 60, 325 Electromotive force (emf). See voltage
Condenser. See capacitor Energy gap, 116
Conduction band, 116 Entanglement, 317
Controlled not, 306 Equivalent digital circuits, 244
Current, 2 Equivalent resistance. See effective
Current divider equation, 7 resistance
Current mirror, 183
Current source
F
ideal, 18
Feedback, amplifier, 190
schematic symbol, 18
FET amplifiers, 145
common drain, 145
common gate, 145
D common source, 145
Darlington pair, 180 Field Effect Transistor (FET), 135
Decade counter, 257 AC model, 142
Decoders and encoders, 246 JFET, 135
Delta-sigma 1-bit D/A, 279 ohmic region, 148
Delta-Y conversion, 40 Filter
DeMorgan’s theorems, 237 LRC, 87
Depletion region, 117 op-amp, 197
Differential amplifier, 181 state variable, 197
Differential amplifier, non-ideal, 210 Flip-flop circuits, 249
Differential voltage gain, 182 binary counter, 257
Digital codes, 265 D flip-flop, 252
Digital logic circuit elements, 237 edge-triggered, 253
Digital to Analog (D/A), 278 gated RS flip-flop, 251
Diode, 115 JK flip-flop, 251
analytic model, 120 Johnson counter, 255
ideal model, 121 ring counter, 255
light emitting (LED), 132 RS flip-flop, 250
models, 119 shift register, 255
p-n junction, 117 Floating point binary numbers, 294
ratings, 129 Fourier series, 322
semiconductor, 115 Fourier transform, 324
specialty, 130 Frequency, 53
turn-on voltage, 119 angular, 53
Diode clamp, 124 Frequency, complex, 99
Index 331

Frequency divider, 255, 257 table, 101


Full adder, 287 LC circuits
Full-wave rectifier, 125 parallel, 85
series, 85
Light Emitting Diode (LED), 132
G Linear device, 51
Gain-bandwidth product, 179 Linear feedback shift register, 256
Gate Linear taper, 23
AND, NAND, 237 Load line
CNOT, 306 BJT, 161
Fredkin, 306 FET, 140
logic, 237 Loop method. See mesh method
NOT, 237
OR, NOR, 237
reversible, 306 M
Toffoli, 306 Maximum forward current, 129
XOR, XNOR, 237 Mesh method, 13
Gray code, 270 Micro-Electromechanical Systems
Ground, 20 (MEMS), 77
Microphone
condenser, 71
H electret, 71
Hadamard gate, 316 Miller’s theorem, 177
Half adder, 287 MOSFET, 148
Half-wave rectifier, 122 depletion, 149
Hexadecimal, 267 enhancement, 149
Hole, 116 Multiplexing, 248
Howland configuration, 202 Mutual inductance, 82
Hybrid parameters, 160

N
I Nodal analysis, 17
Ideal diode model, 121 Node, 4
Ideal transformer, 83 Normalization, 314
IGFET, 149 Norton equivalent, 29
Impedance, complex, 62 Null measurement, 35
Impedance match, 94, 186
Inductors
parallel and series, 63 O
Octal, 267
Ohm’s law, 4
J Ohms per square, 45
Johnson counter, 255 Op-amp, 189
ideal, 189
rail-to-rail, 221
K Op-amp circuits
Kelvin bridge, 42 absolute value, 208, 222
Kirchhoff’s laws buffer, 191
current law (KCL), 5 capacitive sensor, 200
voltage law (KVL), 5 constant current source, 202
difference amplifier, 194
ideal diode, 205
L instrumentation amplifier, 199
Laplace transform, 99 integrator, 196
332 Index

inverting amplifier, 192 Quantization noise, 284


log amplifier, 207 Quantum measurement, 318
low-pass filter, 197 Quantum physics, 311
more output power, 209 Qubit, 315
negative resistance, 202
non-inverting amplifier, 193
peak follower, 206 R
summing amplifier, 194 Reciprocity theorem, 36
twin-T oscillator, 214 Recursive numerical computations, 295
Operating point, 138 Resistance, 3
Operational amplifier. See Op-amp Resistance measurement
Oscillations, amplifiers, 213 four wire, 42
Oscillators, 260 van der Pauw technique, 44
Resistors, 5
in parallel, 7
P in series, 5
Parity, 242 Reverse saturation current, 120
Partial fractions, 103, 110 Rheostat, 23
Passive devices Ring counter, 255
capacitor, 53 Ring oscillator, 260
inductor, 53 RLC circuits, 85
Peak Inverse Voltage (PIV), 129 Root Mean Square (RMS), 53
Period, 53 R-2R ladder, 37, 280
Phase sensitive detector, 232
Phase shift, 58
Photodiode, 133 S
Photons, 311 Saturation region
Phototransistor, 254 BJT, 156
Polarizer, 312 Scanning keyboard, 258
Poles, 111 Schematic
Position sensor diagram, 3
capacitive, 77 symbol, 3
inductive, 85 Schematic symbol
Potentiometer, 23 amplifier, op-amp, 189
Kelvin-Varley, 24 battery, 4
Power bipolar junction transistor, 155
AC, 69 capacitor, 53
electrical, 22 common, 20
RMS, 69 digital gates, 237
Power factor, 70 diode, 118
Power sources, idealized, 54 edge triggered flip-flop, 253
Power Supply Rejection Ratio (PSRR), 183 ground, 20
Priority encoder, 248 ideal sources, 54
Propagation delay, 259 inductor, 53
Pseudo-random sequence, 256 JFET, 135
Pulse density encoding, 279 MOSFETs, 149
Pulse length encoding, 279 resistor, 4
Pulse width encoding, 279 SCR, 184
Push-pull amplifier, 181 switch, 4
triac, 184
Self inductance, 82
Q Semiconductor, 116
Quality factor (Q), 75 doping, n- and p-, 117
Index 333

electrons and holes, 116 equivalent impedance, 90


intrinsic, 117 impedance of finite, 92
Seven-segment display, 247 wave speed, 92
Shift register, 255 Triac, 184
Silicon Controlled Rectifier (SCR), 184 Tri-state, 299
threshold voltage, 184 Truth table, 236
Smith chart, 93 solving circuits with, 241
Standard International (SI) units, 321 Turn-on voltage, diode, 119
State variable filter, 197 Two’s complement, 287
Summing D/A, 280
Superposition theorem, 21
Switch bounce, 226 U
Switch debouncer, 226 Unit prefixes, 321
Switched capacitor methods, 79 Unity gain bandwidth, 179
using tri-state, 300

V
T
Valence band, 116
Tensor notation, 308
Thevenin equivalent, 29 Voltage, 2
Thyristors. See silicon controlled rectifier Voltage divider equation, 5
Time constant, 56 Voltage drop, 4
L/R, 56 Voltmeter, ideal, 7
RC, 56
Transconductance
FET model, 142 W
Transconductance amplifier, 215 Wheatstone bridge, 34
Transfer characteristic, 136, 157
Transfer function, 111
Transformers, 82 Z
dot convention, 84 Zener diodes, 131
Transmission lines, 88 Zeros, 111

You might also like