Anda di halaman 1dari 464

Fifth Edition, last update July 02, 2007

Lessons In Electric Circuits, Volume III Semiconductors


By Tony R. Kuphaldt
Fifth Edition, last update July 02, 2007

i
c
2000-2007,
Tony R. Kuphaldt
This book is published under the terms and conditions of the Design Science License. These
terms and conditions allow for free copying, distribution, and/or modification of this document by
the general public. The full Design Science License text is included in the last chapter.
As an open and collaboratively developed text, this book is distributed in the hope that it
will be useful, but WITHOUT ANY WARRANTY; without even the implied warranty of MERCHANTABILITY or FITNESS FOR A PARTICULAR PURPOSE. See the Design Science License
for more details.
Available in its entirety as part of the Open Book Project collection at:
www.ibiblio.org/obp/electricCircuits

PRINTING HISTORY
First Edition: Printed in June of 2000. Plain-ASCII illustrations for universal computer
readability.
Second Edition: Printed in September of 2000. Illustrations reworked in standard graphic
(eps and jpeg) format. Source files translated to Texinfo format for easy online and printed
publication.
Third Edition: Printed in January 2002. Source files translated to SubML format. SubML is
a simple markup language designed to easily convert to other markups like LATEX, HTML, or
DocBook using nothing but search-and-replace substitutions.
Fourth Edition: Printed in December 2002. New sections added, and error corrections made,
since third edition.
Fith Edition: Printed in July 2007. New sections added, and error corrections made, format
change.

ii

Contents
1 AMPLIFIERS AND ACTIVE DEVICES
1.1 From electric to electronic . . . . . . . .
1.2 Active versus passive devices . . . . . . .
1.3 Amplifiers . . . . . . . . . . . . . . . . .
1.4 Amplifier gain . . . . . . . . . . . . . . .
1.5 Decibels . . . . . . . . . . . . . . . . . . .
1.6 Absolute dB scales . . . . . . . . . . . . .
1.7 Attenuators . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

1
1
3
3
6
7
14
15

2 SOLID-STATE DEVICE THEORY


2.1 Introduction . . . . . . . . . . . . . . . . . . . .
2.2 Quantum physics . . . . . . . . . . . . . . . . .
2.3 Valence and Crystal structure . . . . . . . . . .
2.4 Band theory of solids . . . . . . . . . . . . . . .
2.5 Electrons and holes . . . . . . . . . . . . . . .
2.6 The P-N junction . . . . . . . . . . . . . . . . .
2.7 Junction diodes . . . . . . . . . . . . . . . . . .
2.8 Bipolar junction transistors . . . . . . . . . . . .
2.9 Junction field-effect transistors . . . . . . . . . .
2.10 Insulated-gate field-effect transistors (MOSFET)
2.11 Thyristors . . . . . . . . . . . . . . . . . . . . .
2.12 Semiconductor manufacturing techniques . . . .
2.13 Superconducting devices . . . . . . . . . . . . .
2.14 Quantum devices . . . . . . . . . . . . . . . . .
2.15 Semiconductor devices in SPICE . . . . . . . . .
Bibliography . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

25
25
26
39
45
47
52
55
57
62
67
70
72
76
79
87
90

3 DIODES AND RECTIFIERS


3.1 Meter check of a diode . . .
3.2 Diode ratings . . . . . . . . .
3.3 Rectifier circuits . . . . . . .
3.4 Peak detector . . . . . . . .
3.5 Clipper circuits . . . . . . .
3.6 Clamper circuits . . . . . . .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

93
99
103
104
111
112
116

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

iii

.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

iv

CONTENTS
3.7 Voltage multipliers . . . . . . .
3.8 Inductor commutating circuits
3.9 Diode switching circuits . . . .
3.10 Zener diodes . . . . . . . . . .
3.11 Special-purpose diodes . . . .
3.12 Other diode technologies . . .
3.13 SPICE models . . . . . . . . .
Bibliography . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

118
124
127
129
137
157
157
165

4 BIPOLAR JUNCTION TRANSISTORS


4.1 Introduction . . . . . . . . . . . . . . . .
4.2 The transistor as a switch . . . . . . . . .
4.3 Meter check of a transistor . . . . . . . .
4.4 Active mode operation . . . . . . . . . .
4.5 The common-emitter amplifier . . . . . .
4.6 The common-collector amplifier . . . . .
4.7 The common-base amplifier . . . . . . . .
4.8 Biasing techniques . . . . . . . . . . . . .
4.9 Input and output coupling . . . . . . . .
4.10 Feedback . . . . . . . . . . . . . . . . . .
4.11 Amplifier impedances . . . . . . . . . . .
4.12 Current mirrors . . . . . . . . . . . . . .
4.13 Transistor ratings and packages . . . . .
4.14 BJT quirks . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

167
167
170
173
178
186
201
209
215
228
235
242
243
246
247

5 JUNCTION FIELD-EFFECT TRANSISTORS


5.1 Introduction . . . . . . . . . . . . . . . . . . . .
5.2 The transistor as a switch . . . . . . . . . . . . .
5.3 Meter check of a transistor . . . . . . . . . . . .
5.4 Active-mode operation . . . . . . . . . . . . . .
5.5 The common-source amplifier PENDING . . .
5.6 The common-drain amplifier PENDING . . .
5.7 The common-gate amplifier PENDING . . . .
5.8 Biasing techniques PENDING . . . . . . . . .
5.9 Transistor ratings and packages PENDING . .
5.10 JFET quirks PENDING . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

249
249
251
254
256
265
266
266
266
266
267

6 INSULATED-GATE FIELD-EFFECT TRANSISTORS


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . .
6.2 Depletion-type IGFETs . . . . . . . . . . . . . . . . . . .
6.3 Enhancement-type IGFETs PENDING . . . . . . . . .
6.4 Active-mode operation PENDING . . . . . . . . . . . .
6.5 The common-source amplifier PENDING . . . . . . . .
6.6 The common-drain amplifier PENDING . . . . . . . .
6.7 The common-gate amplifier PENDING . . . . . . . . .
6.8 Biasing techniques PENDING . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

269
269
270
279
279
280
280
280
280

CONTENTS
6.9
6.10
6.11
6.12

Transistor ratings and packages PENDING


IGFET quirks PENDING . . . . . . . . . .
MESFETs PENDING . . . . . . . . . . . .
IGBTs . . . . . . . . . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

280
281
281
281

7 THYRISTORS
7.1 Hysteresis . . . . . . . . . . . . . . . . . . . . . .
7.2 Gas discharge tubes . . . . . . . . . . . . . . . .
7.3 The Shockley Diode . . . . . . . . . . . . . . . .
7.4 The DIAC . . . . . . . . . . . . . . . . . . . . .
7.5 The Silicon-Controlled Rectifier (SCR) . . . . .
7.6 The TRIAC . . . . . . . . . . . . . . . . . . . .
7.7 Optothyristors . . . . . . . . . . . . . . . . . . .
7.8 The Unijunction Transistor (UJT) PENDING
7.9 The Silicon-Controlled Switch (SCS) . . . . . . .
7.10 Field-effect-controlled thyristors . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

285
285
286
290
296
297
308
310
311
311
313

8 OPERATIONAL AMPLIFIERS
8.1 Introduction . . . . . . . . . . . . . . .
8.2 Single-ended and differential amplifiers
8.3 The operational amplifier . . . . . . .
8.4 Negative feedback . . . . . . . . . . . .
8.5 Divided feedback . . . . . . . . . . . . .
8.6 An analogy for divided feedback . . . .
8.7 Voltage-to-current signal conversion . .
8.8 Averager and summer circuits . . . . .
8.9 Building a differential amplifier . . . . .
8.10 The instrumentation amplifier . . . . .
8.11 Differentiator and integrator circuits . .
8.12 Positive feedback . . . . . . . . . . . . .
8.13 Practical considerations . . . . . . . . .
8.14 Operational amplifier models . . . . . .
8.15 Data . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

315
315
316
320
326
329
332
337
338
341
343
344
347
351
365
371

9 PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS


9.1 Power supply circuits INCOMPLETE . . . . . . . . . . . .
9.2 Amplifier circuits PENDING . . . . . . . . . . . . . . . . .
9.3 Oscillator circuits INCOMPLETE . . . . . . . . . . . . . .
9.4 Phase-locked loops PENDING . . . . . . . . . . . . . . . .
9.5 Radio circuits INCOMPLETE . . . . . . . . . . . . . . . .
9.6 Computational circuits . . . . . . . . . . . . . . . . . . . . .
9.7 Measurement circuits INCOMPLETE . . . . . . . . . . . .
9.8 Control circuits PENDING . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

373
373
375
375
376
376
379
400
401

10 ACTIVE FILTERS

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

403

vi

CONTENTS

11 DC MOTOR DRIVES

405

12 INVERTERS AND AC MOTOR DRIVES

407

13 ELECTRON TUBES
13.1 Introduction . . . . . . . . .
13.2 Early tube history . . . . . .
13.3 The triode . . . . . . . . . .
13.4 The tetrode . . . . . . . . . .
13.5 Beam power tubes . . . . . .
13.6 The pentode . . . . . . . . .
13.7 Combination tubes . . . . .
13.8 Tube parameters . . . . . . .
13.9 Ionization (gas-filled) tubes .
13.10 Display tubes . . . . . . . . .
13.11 Microwave tubes . . . . . . .
13.12 Tubes versus Semiconductors

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

409
409
410
413
415
416
417
418
421
423
427
430
433

A-1 ABOUT THIS BOOK

437

A-2 CONTRIBUTOR LIST

441

A-3 DESIGN SCIENCE LICENSE

445

INDEX

449

Chapter 1

AMPLIFIERS AND ACTIVE


DEVICES
Contents
1.1
1.2
1.3
1.4
1.5
1.6
1.7

1.1

From electric to electronic . . . . . . . . . .


Active versus passive devices . . . . . . . . .
Amplifiers . . . . . . . . . . . . . . . . . . . .
Amplifier gain . . . . . . . . . . . . . . . . . .
Decibels . . . . . . . . . . . . . . . . . . . . . .
Absolute dB scales . . . . . . . . . . . . . . .
Attenuators . . . . . . . . . . . . . . . . . . .
1.7.1
Decibels . . . . . . . . . . . . . . . . . . . .
1.7.2
T-section attenuator . . . . . . . . . . . . . .
1.7.3
-section attenuator . . . . . . . . . . . . .
1.7.4
L-section attenuator . . . . . . . . . . . . . .
1.7.5
Bridged T attenuator . . . . . . . . . . . . .
1.7.6
Cascaded sections . . . . . . . . . . . . . . .
1.7.7
RF attenuators . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .

1
3
3
6
7
14
15
16
18
19
20
20
20
22

From electric to electronic

This third volume of the book series Lessons In Electric Circuits makes a departure from the former
two in that the transition between electric circuits and electronic circuits is formally crossed. Electric
circuits are connections of conductive wires and other devices whereby the uniform flow of electrons
occurs. Electronic circuits add a new dimension to electric circuits in that some means of control is
exerted over the flow of electrons by another electrical signal, either a voltage or a current.
In and of itself, the control of electron flow is nothing new to the student of electric circuits.
Switches control the flow of electrons, as do potentiometers, especially when connected as variable
1

CHAPTER 1. AMPLIFIERS AND ACTIVE DEVICES

resistors (rheostats). Neither the switch nor the potentiometer should be new to your experience
by this point in your study. The threshold marking the transition from electric to electronic, then,
is defined by how the flow of electrons is controlled rather than whether or not any form of control
exists in a circuit. Switches and rheostats control the flow of electrons according to the positioning of
a mechanical device, which is actuated by some physical force external to the circuit. In electronics,
however, we are dealing with special devices able to control the flow of electrons according to another
flow of electrons, or by the application of a static voltage. In other words, in an electronic circuit,
electricity is able to control electricity.
The historic precursor to the modern electronics era was invented by Thomas Edison in 1880
while developing the electric incandescent lamp. Edison found that a small current passed from the
heated lamp filament to a metal plate mounted inside the vacuum envelop. (Figure 1.1 (a)) Today
this is known as the Edison effect. Note that the battery is only necessary to heat the filament.
Electrons would still flow if a non-electrical heat source was used.
control
e-1

e-1

(a)

(b)

e-1

- +
(c)

Figure 1.1: (a) Edison effect, (b) Fleming valve or vacuum diode, (c) DeForest audion triode vacuum
tube amplifier.
By 1904 Marconi wireless company adviser John Flemming found that an externally applied
current (plate battery) only passed in one direction from filament to plate (Figure 1.1 (b)), but not
the reverse direction (not shown). This invention was the vacuum diode, used to convert alternating
currents to DC. The addition of a third electrode by Lee DeForest (Figure 1.1 (c)) allowed a small
signal to control the larger electron flow from filament to plate.
Historically, the era of electronics began with the invention of the Audion tube, a device controlling
the flow of an electron stream through a vacuum by the application of a small voltage between two
metal structures within the tube. A more detailed summary of so-called electron tube or vacuum
tube technology is available in the last chapter of this volume for those who are interested.
Electronics technology experienced a revolution in 1948 with the invention of the transistor.
This tiny device achieved approximately the same effect as the Audion tube, but in a vastly smaller
amount of space and with less material. Transistors control the flow of electrons through solid
semiconductor substances rather than through a vacuum, and so transistor technology is often
referred to as solid-state electronics.

1.2.

1.2

ACTIVE VERSUS PASSIVE DEVICES

Active versus passive devices

An active device is any type of circuit component with the ability to electrically control electron
flow (electricity controlling electricity). In order for a circuit to be properly called electronic, it must
contain at least one active device. Components incapable of controlling current by means of another
electrical signal are called passive devices. Resistors, capacitors, inductors, transformers, and even
diodes are all considered passive devices. Active devices include, but are not limited to, vacuum
tubes, transistors, silicon-controlled rectifiers (SCRs), and TRIACs. A case might be made for the
saturable reactor to be defined as an active device, since it is able to control an AC current with a
DC current, but Ive never heard it referred to as such. The operation of each of these active devices
will be explored in later chapters of this volume.
All active devices control the flow of electrons through them. Some active devices allow a voltage
to control this current while other active devices allow another current to do the job. Devices utilizing
a static voltage as the controlling signal are, not surprisingly, called voltage-controlled devices.
Devices working on the principle of one current controlling another current are known as currentcontrolled devices. For the record, vacuum tubes are voltage-controlled devices while transistors are
made as either voltage-controlled or current controlled types. The first type of transistor successfully
demonstrated was a current-controlled device.

1.3

Amplifiers

The practical benefit of active devices is their amplifying ability. Whether the device in question
be voltage-controlled or current-controlled, the amount of power required of the controlling signal
is typically far less than the amount of power available in the controlled current. In other words,
an active device doesnt just allow electricity to control electricity; it allows a small amount of
electricity to control a large amount of electricity.
Because of this disparity between controlling and controlled powers, active devices may be employed to govern a large amount of power (controlled) by the application of a small amount of power
(controlling). This behavior is known as amplification.
It is a fundamental rule of physics that energy can neither be created nor destroyed. Stated
formally, this rule is known as the Law of Conservation of Energy, and no exceptions to it have been
discovered to date. If this Law is true and an overwhelming mass of experimental data suggests
that it is then it is impossible to build a device capable of taking a small amount of energy and
magically transforming it into a large amount of energy. All machines, electric and electronic circuits
included, have an upper efficiency limit of 100 percent. At best, power out equals power in as in
Figure 1.2.
Usually, machines fail even to meet this limit, losing some of their input energy in the form of
heat which is radiated into surrounding space and therefore not part of the output energy stream.
(Figure 1.3)
Many people have attempted, without success, to design and build machines that output more
power than they take in. Not only would such a perpetual motion machine prove that the Law of
Conservation of Energy was not a Law after all, but it would usher in a technological revolution such
as the world has never seen, for it could power itself in a circular loop and generate excess power
for free. (Figure 1.4)
Despite much effort and many unscrupulous claims of free energy or over-unity machines, not

CHAPTER 1. AMPLIFIERS AND ACTIVE DEVICES

Perfect machine

Pinput

Poutput

Efficiency =

Poutput

= 1 = 100%

Pinput

Figure 1.2: The power output of a machine can approach, but never exceed, the power input for
100% efficiency as an upper limit.

Realistic machine

Pinput

Poutput

Plost (usually waste heat)

Efficiency =

Poutput
Pinput

< 1 = less than 100%

Figure 1.3: A realistic machine most often looses some of its input energy as heat in transforming
it into the output energy stream.

Pinput

Perpetual-motion
machine

Efficiency =

Pinput

Poutput
Pinput

Poutput

> 1 = more than 100%

Perpetual-motion
machine

P"free"
Poutput

Figure 1.4: Hypothetical perpetual motion machine powers itself?

1.3.

AMPLIFIERS

one has ever passed the simple test of powering itself with its own energy output and generating
energy to spare.
There does exist, however, a class of machines known as amplifiers, which are able to take in
small-power signals and output signals of much greater power. The key to understanding how
amplifiers can exist without violating the Law of Conservation of Energy lies in the behavior of
active devices.
Because active devices have the ability to control a large amount of electrical power with a small
amount of electrical power, they may be arranged in circuit so as to duplicate the form of the input
signal power from a larger amount of power supplied by an external power source. The result is a
device that appears to magically magnify the power of a small electrical signal (usually an AC voltage
waveform) into an identically-shaped waveform of larger magnitude. The Law of Conservation of
Energy is not violated because the additional power is supplied by an external source, usually a DC
battery or equivalent. The amplifier neither creates nor destroys energy, but merely reshapes it into
the waveform desired as shown in Figure 1.5.

External
power source

Pinput

Amplifier

Poutput

Figure 1.5: While an amplifier can scale a small input signal to large output, its energy source is an
external power supply.
In other words, the current-controlling behavior of active devices is employed to shape DC power
from the external power source into the same waveform as the input signal, producing an output
signal of like shape but different (greater) power magnitude. The transistor or other active device
within an amplifier merely forms a larger copy of the input signal waveform out of the raw DC
power provided by a battery or other power source.
Amplifiers, like all machines, are limited in efficiency to a maximum of 100 percent. Usually,
electronic amplifiers are far less efficient than that, dissipating considerable amounts of energy in
the form of waste heat. Because the efficiency of an amplifier is always 100 percent or less, one can
never be made to function as a perpetual motion device.
The requirement of an external source of power is common to all types of amplifiers, electrical
and non-electrical. A common example of a non-electrical amplification system would be power
steering in an automobile, amplifying the power of the drivers arms in turning the steering wheel
to move the front wheels of the car. The source of power necessary for the amplification comes from
the engine. The active device controlling the drivers input signal is a hydraulic valve shuttling
fluid power from a pump attached to the engine to a hydraulic piston assisting wheel motion. If the

CHAPTER 1. AMPLIFIERS AND ACTIVE DEVICES

engine stops running, the amplification system fails to amplify the drivers arm power and the car
becomes very difficult to turn.

1.4

Amplifier gain

Because amplifiers have the ability to increase the magnitude of an input signal, it is useful to be
able to rate an amplifiers amplifying ability in terms of an output/input ratio. The technical term
for an amplifiers output/input magnitude ratio is gain. As a ratio of equal units (power out / power
in, voltage out / voltage in, or current out / current in), gain is naturally a unitless measurement.
Mathematically, gain is symbolized by the capital letter A.
For example, if an amplifier takes in an AC voltage signal measuring 2 volts RMS and outputs
an AC voltage of 30 volts RMS, it has an AC voltage gain of 30 divided by 2, or 15:

AV =
AV =

Voutput
Vinput
30 V
2V

AV = 15
Correspondingly, if we know the gain of an amplifier and the magnitude of the input signal, we
can calculate the magnitude of the output. For example, if an amplifier with an AC current gain of
3.5 is given an AC input signal of 28 mA RMS, the output will be 3.5 times 28 mA, or 98 mA:

Ioutput = (AI)(Iinput)
Ioutput = (3.5)(28 mA)
Ioutput = 98 mA
In the last two examples I specifically identified the gains and signal magnitudes in terms of
AC. This was intentional, and illustrates an important concept: electronic amplifiers often respond
differently to AC and DC input signals, and may amplify them to different extents. Another way
of saying this is that amplifiers often amplify changes or variations in input signal magnitude (AC)
at a different ratio than steady input signal magnitudes (DC). The specific reasons for this are too
complex to explain at this time, but the fact of the matter is worth mentioning. If gain calculations
are to be carried out, it must first be understood what type of signals and gains are being dealt
with, AC or DC.
Electrical amplifier gains may be expressed in terms of voltage, current, and/or power, in both
AC and DC. A summary of gain definitions is as follows. The triangle-shaped delta symbol ()
represents change in mathematics, so Voutput / Vinput means change in output voltage divided
by change in input voltage, or more simply, AC output voltage divided by AC input voltage:

1.5.

DECIBELS

DC gains
Voltage

AV =

Current

AI =

AC gains

Voutput

AV =

Vinput
Ioutput

AI =

Iinput
Poutput

Power

AP =

Pinput

AP =

Voutput
Vinput
Ioutput
Iinput

(Voutput)(Ioutput)
(Vinput)(Iinput)

AP = (AV)(AI)

= "change in . . ."
If multiple amplifiers are staged, their respective gains form an overall gain equal to the product
(multiplication) of the individual gains. (Figure 1.6) If a 1 V signal were applied to the input of the
gain of 3 amplifier in Figure 1.6 a 3 V signal out of the first amplifier would be further amplified by
a gain of 5 at the second stage yielding 15 V at the final output.
Input signal

Amplifier
gain = 3

Amplifier
gain = 5

Output signal

Overall gain = (3)(5) = 15

Figure 1.6: The gain of a chain of cascaded amplifiers is the product of the individual gains.

1.5

Decibels

In its simplest form, an amplifiers gain is a ratio of output over input. Like all ratios, this form of
gain is unitless. However, there is an actual unit intended to represent gain, and it is called the bel.
As a unit, the bel was actually devised as a convenient way to represent power loss in telephone
system wiring rather than gain in amplifiers. The units name is derived from Alexander Graham
Bell, the famous Scottish inventor whose work was instrumental in developing telephone systems.
Originally, the bel represented the amount of signal power loss due to resistance over a standard
length of electrical cable. Now, it is defined in terms of the common (base 10) logarithm of a power
ratio (output power divided by input power):

AP(ratio) =

Poutput
Pinput

AP(Bel) = log

Poutput
Pinput

CHAPTER 1. AMPLIFIERS AND ACTIVE DEVICES

Because the bel is a logarithmic unit, it is nonlinear. To give you an idea of how this works,
consider the following table of figures, comparing power losses and gains in bels versus simple ratios:

Table: Gain / loss in bels


Loss/gain as
a ratio
Poutput
Pinput

Loss/gain
in bels
Poutput
log
Pinput

Loss/gain as
a ratio
Poutput
Pinput

Loss/gain
in bels
Poutput
log
Pinput

1000

3B

0.1

-1 B

100

2B

0.01

-2 B

10

1B

0.001

-3 B

0B

0.0001

-4 B

1
(no loss or gain)

It was later decided that the bel was too large of a unit to be used directly, and so it became
customary to apply the metric prefix deci (meaning 1/10) to it, making it deci bels, or dB. Now, the
expression dB is so common that many people do not realize it is a combination of deci- and
-bel, or that there even is such a unit as the bel. To put this into perspective, here is another
table contrasting power gain/loss ratios against decibels:

Table: Gain / loss in decibels


Loss/gain as
a ratio
Poutput
Pinput

Loss/gain
in decibels
Poutput
10 log
Pinput

Loss/gain as
a ratio
Poutput
Pinput

Loss/gain
in decibels
Poutput
10 log
Pinput

1000

30 dB

0.1

-10 dB

100

20 dB

0.01

-20 dB

10

10 dB

0.001

-30 dB

0 dB

0.0001

-40 dB

1
(no loss or gain)

As a logarithmic unit, this mode of power gain expression covers a wide range of ratios with a
minimal span in figures. It is reasonable to ask, why did anyone feel the need to invent a logarithmic
unit for electrical signal power loss in a telephone system? The answer is related to the dynamics
of human hearing, the perceptive intensity of which is logarithmic in nature.
Human hearing is highly nonlinear: in order to double the perceived intensity of a sound, the
actual sound power must be multiplied by a factor of ten. Relating telephone signal power loss

1.5.

DECIBELS

in terms of the logarithmic bel scale makes perfect sense in this context: a power loss of 1 bel
translates to a perceived sound loss of 50 percent, or 1/2. A power gain of 1 bel translates to a
doubling in the perceived intensity of the sound.
An almost perfect analogy to the bel scale is the Richter scale used to describe earthquake
intensity: a 6.0 Richter earthquake is 10 times more powerful than a 5.0 Richter earthquake; a 7.0
Richter earthquake 100 times more powerful than a 5.0 Richter earthquake; a 4.0 Richter earthquake
is 1/10 as powerful as a 5.0 Richter earthquake, and so on. The measurement scale for chemical pH
is likewise logarithmic, a difference of 1 on the scale is equivalent to a tenfold difference in hydrogen
ion concentration of a chemical solution. An advantage of using a logarithmic measurement scale is
the tremendous range of expression afforded by a relatively small span of numerical values, and it is
this advantage which secures the use of Richter numbers for earthquakes and pH for hydrogen ion
activity.
Another reason for the adoption of the bel as a unit for gain is for simple expression of system
gains and losses. Consider the last system example (Figure 1.6) where two amplifiers were connected
tandem to amplify a signal. The respective gain for each amplifier was expressed as a ratio, and the
overall gain for the system was the product (multiplication) of those two ratios:
Overall gain = (3)(5) = 15
If these figures represented power gains, we could directly apply the unit of bels to the task of
representing the gain of each amplifier, and of the system altogether. (Figure 1.7)
AP(Bel) = log AP(ratio)

Input signal

AP(Bel) = log 3

AP(Bel) = log 5

Amplifier
gain = 3
gain = 0.477 B

Amplifier
gain = 5
gain = 0.699 B

Output signal

Overall gain = (3)(5) = 15


Overall gain(Bel) = log 15 = 1.176 B

Figure 1.7: Power gain in bels is additive: 0.477 B + 0.699 B = 1.176 B.


Close inspection of these gain figures in the unit of bel yields a discovery: theyre additive.
Ratio gain figures are multiplicative for staged amplifiers, but gains expressed in bels add rather
than multiply to equal the overall system gain. The first amplifier with its power gain of 0.477 B
adds to the second amplifiers power gain of 0.699 B to make a system with an overall power gain
of 1.176 B.
Recalculating for decibels rather than bels, we notice the same phenomenon. (Figure 1.8)
To those already familiar with the arithmetic properties of logarithms, this is no surprise. It is an
elementary rule of algebra that the antilogarithm of the sum of two numbers logarithm values equals
the product of the two original numbers. In other words, if we take two numbers and determine the
logarithm of each, then add those two logarithm figures together, then determine the antilogarithm
of that sum (elevate the base number of the logarithm in this case, 10 to the power of that sum),
the result will be the same as if we had simply multiplied the two original numbers together. This
algebraic rule forms the heart of a device called a slide rule, an analog computer which could, among

10

CHAPTER 1. AMPLIFIERS AND ACTIVE DEVICES


AP(dB) = 10 log AP(ratio)

Input signal

AP(dB) = 10 log 3

AP(dB) = 10 log 5

Amplifier
gain = 3
gain = 4.77 dB

Amplifier
gain = 5
gain = 6.99 dB

Output signal

Overall gain = (3)(5) = 15


Overall gain(dB) = 10 log 15 = 11.76 dB

Figure 1.8: Gain of amplifier stages in decibels is additive: 4.77 dB + 6.99 dB = 11.76 dB.
other things, determine the products and quotients of numbers by addition (adding together physical
lengths marked on sliding wood, metal, or plastic scales). Given a table of logarithm figures, the
same mathematical trick could be used to perform otherwise complex multiplications and divisions
by only having to do additions and subtractions, respectively. With the advent of high-speed,
handheld, digital calculator devices, this elegant calculation technique virtually disappeared from
popular use. However, it is still important to understand when working with measurement scales
that are logarithmic in nature, such as the bel (decibel) and Richter scales.
When converting a power gain from units of bels or decibels to a unitless ratio, the mathematical
inverse function of common logarithms is used: powers of 10, or the antilog.

If:
AP(Bel) = log AP(ratio)
Then:
AP(ratio) = 10AP(Bel)
Converting decibels into unitless ratios for power gain is much the same, only a division factor
of 10 is included in the exponent term:

If:
AP(dB) = 10 log AP(ratio)
Then:

AP(dB)

AP(ratio) = 10

10

Example: Power into an amplifier is 1 watts, the power out is 10 watt. Find the power gain
in dB.
AP (dB) = 10 log10 (PO / PI ) = 10 log10 (10 /1) = 10 log10 (10) = 10 (1) = 10 dB
Example: Find the power gain ratio AP (ratio) = (PO / PI ) for a 20 dB Power gain.
AP (dB) = 20 = 10 log10 AP (ratio)
20/10 = log10 AP (ratio)

1.5.

DECIBELS

11

1020/10 = 10log10 (AP (ratio) )


100 = AP (ratio) = (PO / PI )
Because the bel is fundamentally a unit of power gain or loss in a system, voltage or current
gains and losses dont convert to bels or dB in quite the same way. When using bels or decibels to
express a gain other than power, be it voltage or current, we must perform the calculation in terms
of how much power gain there would be for that amount of voltage or current gain. For a constant
load impedance, a voltage or current gain of 2 equates to a power gain of 4 (2 2 ); a voltage or current
gain of 3 equates to a power gain of 9 (32 ). If we multiply either voltage or current by a given factor,
then the power gain incurred by that multiplication will be the square of that factor. This relates
back to the forms of Joules Law where power was calculated from either voltage or current, and
resistance:

P=

E2
R

P = I2R
Power is proportional to the square
of either voltage or current
Thus, when translating a voltage or current gain ratio into a respective gain in terms of the bel
unit, we must include this exponent in the equation(s):

AP(Bel) = log AP(ratio)


AV(Bel) = log AV(ratio)2

Exponent required

AI(Bel) = log AI(ratio)2


The same exponent requirement holds true when expressing voltage or current gains in terms of
decibels:

AP(dB) = 10 log AP(ratio)


AV(dB) = 10 log AV(ratio)2

Exponent required

AI(dB) = 10 log AI(ratio)2


However, thanks to another interesting property of logarithms, we can simplify these equations
to eliminate the exponent by including the 2 as a multiplying factor for the logarithm function.
In other words, instead of taking the logarithm of the square of the voltage or current gain, we just
multiply the voltage or current gains logarithm figure by 2 and the final result in bels or decibels
will be the same:

12

CHAPTER 1. AMPLIFIERS AND ACTIVE DEVICES

For bels:
AV(Bel) = log AV(ratio)2
. . . is the same as . . .
AV(Bel) = 2 log AV(ratio)

AI(Bel) = log AI(ratio)2


. . . is the same as . . .
AI(Bel) = 2 log AI(ratio)

For decibels:
AV(dB) = 10 log AV(ratio)2
. . . is the same as . . .
AV(dB) = 20 log AV(ratio)

AI(dB) = 10 log AI(ratio)2


. . . is the same as . . .
AI(dB) = 20 log AI(ratio)

The process of converting voltage or current gains from bels or decibels into unitless ratios is
much the same as it is for power gains:

If:
AV(Bel) = 2 log AV(ratio)
Then:

AI(Bel) = 2 log AI(ratio)


AI(Bel)

AV(Bel)

AV(ratio) = 10

AI(ratio) = 10

Here are the equations used for converting voltage or current gains in decibels into unitless ratios:

If:
AV(dB) = 20 log AV(ratio)
Then:

AI(dB) = 20 log AI(ratio)

AV(dB)

AV(ratio) = 10 20

AI(dB)

AI(ratio) = 10 20

While the bel is a unit naturally scaled for power, another logarithmic unit has been invented to
directly express voltage or current gains/losses, and it is based on the natural logarithm rather than
the common logarithm as bels and decibels are. Called the neper, its unit symbol is a lower-case
n.

AV(ratio) =

Voutput
Vinput

AV(neper) = ln AV(ratio)

AI(ratio) =

Ioutput
Iinput

AI(neper) = ln AI(ratio)

For better or for worse, neither the neper nor its attenuated cousin, the decineper, is popularly
used as a unit in American engineering applications.
Example: The voltage into a 600 audio line amplifier is 10 mV, the voltage across a 600
load is 1 V. Find the power gain in dB.
A(dB) = 20 log10 (VO / VI ) = 20 log10 (1 /0.01) = 20 log10 (100) = 20 (2) = 40 dB

1.5.

DECIBELS

13

Example: Find the voltage gain ratio AV (ratio) = (VO / VI ) for a 20 dB gain amplifier having
a 50 input and out impedance.
AV (dB) = 20 log10 AV (ratio)
20 = 20 log10 AV (ratio)
20/20 = log10 AP (ratio)
1020/20 = 10log10 (AV (ratio) )
10 = AV (ratio) = (VO / VI )
REVIEW:
Gains and losses may be expressed in terms of a unitless ratio, or in the unit of bels (B) or
decibels (dB). A decibel is literally a deci -bel: one-tenth of a bel.
The bel is fundamentally a unit for expressing power gain or loss. To convert a power ratio to
either bels or decibels, use one of these equations:

AP(Bel) = log AP(ratio)

AP(db) = 10 log AP(ratio)

When using the unit of the bel or decibel to express a voltage or current ratio, it must be cast
in terms of an equivalent power ratio. Practically, this means the use of different equations,
with a multiplication factor of 2 for the logarithm value corresponding to an exponent of 2 for
the voltage or current gain ratio:

AV(Bel) = 2 log AV(ratio)

AV(dB) = 20 log AV(ratio)

AI(Bel) = 2 log AI(ratio)

AI(dB) = 20 log AI(ratio)

To convert a decibel gain into a unitless ratio gain, use one of these equations:
AV(dB)

AV(ratio) = 10 20
AI(dB)
20

AI(ratio) = 10

AP(dB)

AP(ratio) = 10

10

A gain (amplification) is expressed as a positive bel or decibel figure. A loss (attenuation) is


expressed as a negative bel or decibel figure. Unity gain (no gain or loss; ratio = 1) is expressed
as zero bels or zero decibels.
When calculating overall gain for an amplifier system composed of multiple amplifier stages,
individual gain ratios are multiplied to find the overall gain ratio. Bel or decibel figures for
each amplifier stage, on the other hand, are added together to determine overall gain.

14

1.6

CHAPTER 1. AMPLIFIERS AND ACTIVE DEVICES

Absolute dB scales

It is also possible to use the decibel as a unit of absolute power, in addition to using it as an expression
of power gain or loss. A common example of this is the use of decibels as a measurement of sound
pressure intensity. In cases like these, the measurement is made in reference to some standardized
power level defined as 0 dB. For measurements of sound pressure, 0 dB is loosely defined as the
lower threshold of human hearing, objectively quantified as 1 picowatt of sound power per square
meter of area.
A sound measuring 40 dB on the decibel sound scale would be 104 times greater than the
threshold of hearing. A 100 dB sound would be 1010 (ten billion) times greater than the threshold
of hearing.
Because the human ear is not equally sensitive to all frequencies of sound, variations of the decibel
sound-power scale have been developed to represent physiologically equivalent sound intensities at
different frequencies. Some sound intensity instruments were equipped with filter networks to give
disproportionate indications across the frequency scale, the intent of which to better represent the
effects of sound on the human body. Three filtered scales became commonly known as the A, B,
and C weighted scales. Decibel sound intensity indications measured through these respective
filtering networks were given in units of dBA, dBB, and dBC. Today, the A-weighted scale is
most commonly used for expressing the equivalent physiological impact on the human body, and is
especially useful for rating dangerously loud noise sources.
Another standard-referenced system of power measurement in the unit of decibels has been
established for use in telecommunications systems. This is called the dBm scale. (Figure 1.9) The
reference point, 0 dBm, is defined as 1 milliwatt of electrical power dissipated by a 600 load.
According to this scale, 10 dBm is equal to 10 times the reference power, or 10 milliwatts; 20 dBm is
equal to 100 times the reference power, or 100 milliwatts. Some AC voltmeters come equipped with a
dBm range or scale (sometimes labeled DB) intended for use in measuring AC signal power across
a 600 load. 0 dBm on this scale is, of course, elevated above zero because it represents something
greater than 0 (actually, it represents 0.7746 volts across a 600 load, voltage being equal to the
square root of power times resistance; the square root of 0.001 multiplied by 600). When viewed
on the face of an analog meter movement, this dBm scale appears compressed on the left side and
expanded on the right in a manner not unlike a resistance scale, owing to its logarithmic nature.
Radio frequency power measurements for low level signals encountered in radio receivers use dBm
measurements referenced to a 50 load. Signal generators for the evaluation of radio receivers may
output an adjustable dBm rated signal. The signal level is selected by a device called an attenuator,
described in the next section.
An adaptation of the dBm scale for audio signal strength is used in studio recording and broadcast
engineering for standardizing volume levels, and is called the VU scale. VU meters are frequently
seen on electronic recording instruments to indicate whether or not the recorded signal exceeds the
maximum signal level limit of the device, where significant distortion will occur. This volume
indicator scale is calibrated in according to the dBm scale, but does not directly indicate dBm for
any signal other than steady sine-wave tones. The proper unit of measurement for a VU meter is
volume units.
When relatively large signals are dealt with, and an absolute dB scale would be useful for representing signal level, specialized decibel scales are sometimes used with reference points greater
than the 1 mW used in dBm. Such is the case for the dBW scale, with a reference point of 0 dBW
established at 1 watt. Another absolute measure of power called the dBk scale references 0 dBk at

1.7.

ATTENUATORS

15

Table: Absolute power levels in dBm (decibel milliwatt)


Power in
watts

Power in
milliwatts

Power in
dBm

Power in
milliwatts

Power in
dBm

1000

30 dB

0 dB

0.1

100

20 dB

0.1

-10 dB

0.01

10

10 dB

0.001

-20 dB

0.004

6 dB

0.0001

-30 dB

0.002

3 dB

0.00001

-40 dB

Figure 1.9: Absolute power levels in dBm (decibels referenced to 1 milliwatt.


1 kW, or 1000 watts.
REVIEW:
The unit of the bel or decibel may also be used to represent an absolute measurement of power
rather than just a relative gain or loss. For sound power measurements, 0 dB is defined as a
standardized reference point of power equal to 1 picowatt per square meter. Another dB scale
suited for sound intensity measurements is normalized to the same physiological effects as a
1000 Hz tone, and is called the dBA scale. In this system, 0 dBA is defined as any frequency
sound having the same physiological equivalence as a 1 picowatt-per-square-meter tone at 1000
Hz.
An electrical dB scale with an absolute reference point has been made for use in telecommunications systems. Called the dBm scale, its reference point of 0 dBm is defined as 1 milliwatt
of AC signal power dissipated by a 600 load.
A VU meter reads audio signal level according to the dBm for sine-wave signals. Because
its response to signals other than steady sine waves is not the same as true dBm, its unit of
measurement is volume units.
dB scales with greater absolute reference points than the dBm scale have been invented for
high-power signals. The dBW scale has its reference point of 0 dBW defined as 1 watt of
power. The dBk scale sets 1 kW (1000 watts) as the zero-point reference.

1.7

Attenuators

Attenuators are passive devices. It is convenient to discuss them along with decibels. Attenuators
weaken or attenuate the high level output of a signal generator, for example, to provide a lower

16

CHAPTER 1. AMPLIFIERS AND ACTIVE DEVICES

level signal for something like the antenna input of a sensitive radio receiver. (Figure 1.10) The
attenuator could be built into the signal generator, or be a stand-alone device. It could provide a
fixed or adjustable amount of attenuation. An attenuator section can also provide isolation between
a source and a troublesome load.

ZI

ZI

Attenuator

ZO

ZO

Figure 1.10: Constant impedance attenuator is matched to source impedance ZI and load impedance
ZO . For radio frequency equipment Z is 50 .
In the case of a stand-alone attenuator, it must be placed in series between the signal source
and the load by breaking open the signal path as shown above. In addition, it must match both the
source impedance ZI and the load impedance ZO , while providing a specified amount of attenuation.
In this section we will only consider the special, and most common, case where the source and load
impedances are equal. Not considered in this section, unequal source and load impedances may be
matched by an attenuator section. However, the formulation is more complex.

T attenuator

attenuator

Figure 1.11: T section and section attenuators are common forms.


Common configurations are the T and networks shown in Figure 1.11 Multiple attenuator
sections may be cascaded when even weaker signals are needed as in Figure 1.19.

1.7.1

Decibels

Voltage ratios, as used in the design of attenuators are often expressed in terms of decibels. The
voltage ratio (K below) must be derived from the attenuation in decibels. Power ratios expressed
as decibels are additive. For example, a 10 dB attenuator followed by a 6 dB attenuator provides
16dB of attenuation overall.
10 dB + 6 db = 16 dB
Changing sound levels are perceptible roughly proportional to the logarithm of the power ratio
(PI / PO ).

1.7.

ATTENUATORS

17

sound level = log10 (PI / PO )


A change of 1 dB in sound level is barely perceptible to a listener, while 2 db is readily perceptible.
An attenuation of 3 dB corresponds to cutting power in half, while a gain of 3 db corresponds to a
doubling of the power level. A gain of -3 dB is the same as an attenuation of +3 dB, corresponding
to half the original power level.
The power change in decibels in terms of power ratio is:
dB = 10 log10 (PI / PO )
Assuming that the load RI at PI is the same as the load resistor RO at PO (RI = RO ), the decibels
may be derived from the voltage ratio (VI / VO ) or current ratio (II / IO ):
PO = V

IO = V O 2 / R = I O 2 R

PI = V I I I = V I 2 / R = I I 2 R
dB = 10 log10 (PI / PO ) = 10 log10 (VI 2 / VO 2 ) = 20 log10 (VI /VO )
dB = 10 log10 (PI / PO ) = 10 log10 (II 2 / IO 2 ) = 20 log10 (II /IO )
The two most often used forms of the decibel equation are:
dB = 10 log10 (PI / PO )

or

dB = 20 log10 (VI / VO )

We will use the latter form, since we need the voltage ratio. Once again, the voltage ratio form
of equation is only applicable where the two corresponding resistors are equal. That is, the source
and load resistance need to be equal.
Example: Power into an attenuator is 10 watts, the power out is 1 watt. Find the attenuation
in dB.
dB = 10 log10 (PI / PO ) = 10 log10 (10 /1) = 10 log10 (10) = 10 (1) = 10 dB
Example: Find the voltage attenuation ratio (K= (VI / VO )) for a 10 dB attenuator.
dB = 10= 20 log10 (VI / VO )
10/20 = log10 (VI / VO )
1010/20 = 10log10 (VI /VO )
3.16 = (VI / VO ) = AP (ratio)
Example: Power into an attenuator is 100 milliwatts, the power out is 1 milliwatt. Find the
attenuation in dB.
dB = 10 log10 (PI / PO ) = 10 log10 (100 /1) = 10 log10 (100) = 10 (2) = 20 dB
Example: Find the voltage attenuation ratio (K= (VI / VO )) for a 20 dB attenuator.
dB = 20= 20 log10 (VI / VO )
1020/20 = 10log10 (VI /VO )
10 = (VI / VO ) = K

18

1.7.2

CHAPTER 1. AMPLIFIERS AND ACTIVE DEVICES

T-section attenuator

The T and attenuators must be connected to a Z source and Z load impedance. The Z-(arrows)
pointing away from the attenuator in the figure below indicate this. The Z-(arrows) pointing toward
the attenuator indicates that the impedance seen looking into the attenuator with a load Z on
the opposite end is Z, Z=50 for our case. This impedance is a constant (50 ) with respect to
attenuation impedance does not change when attenuation is changed.
The table in Figure 1.12 lists resistor values for the T and attenuators to match a 50 source/
load, as is the usual requirement in radio frequency work.
Telephone utility and other audio work often requires matching to 600 . Multiply all R values
by the ratio (600/50) to correct for 600 matching. Multiplying by 75/50 would convert table
values to match a 75 source and load.

dB = attenuation in decibels
Z = source/load impedance (resistive)
K>1
R1
R1
VI
K =
= 10 dB/20
VO
VI
VO
K- 1
R2
R1 = Z

K+ 1
2K
R2 = Z
T attenuator
K2- 1

Resistors for
Z = 50
Attenuation
dB
K=Vi/Vo
1.0
1.12
2.0
1.26
3.0
1.41
4.0
1.58
6.0
2.00
10.0
3.16
20.0
10.00

T-section
R1
2.88
5.73
8.55
11.31
16.61
25.97
40.91

R2
433.34
215.24
141.93
104.83
66.93
35.14
10.10

Figure 1.12: Formulas for T-section attenuator resistors, given K, the voltage attenuation ratio, and
ZI = ZO = 50 .
The amount of attenuation is customarily specified in dB (decibels). Though, we need the voltage
(or current) ratio K to find the resistor values from equations. See the dB/20 term in the power of
10 term for computing the voltage ratio K from dB, above.
The T (and below ) configurations are most commonly used as they provide bidirectional
matching. That is, the attenuator input and output may be swapped end for end and still match
the source and load impedances while supplying the same attenuation.
Disconnecting the source and looking in to the right at VI , we need to see a series parallel
combination of RO , RI , RO , and Z looking like an equivalent resistance of Z, the source/load
impedance: (Z is connected to the output.)
Z=RO + (R2 (R1 + Z))
For example, substitute the 10 dB values from the 50 attenuator table for R 1 and R2 as shown
in Figure 1.13.
Z = 25.97 + (35.14 (25.97 + 50))
Z = 25.97 + (35.14 75.97 )
Z = 25.97 + 24.03 = 50

1.7.

ATTENUATORS

19

This shows us that we see 50 looking right into the example attenuator (Figure 1.13) with a
50 load.
Replacing the source generator, disconnecting load Z at VO , and looking in to the left, should
give us the same equation as above for the impedance at VO , due to symmetry. Moreover, the
three resistors must be values which supply the required attenuation from input to output. This is
accomplished by the equations for R1 and R2 above as applied to the T-attenuator below.

R1=26.0

VI

=50

R1
R2= VO
35.1
=50

T attenuator
10 dB attenuators for matching input/output to Z= 50 .
Figure 1.13: 10 dB T-section attenuator for insertion between a 50 source and load.

1.7.3

-section attenuator

The table in Figure 1.14 lists resistor values for the attenuator matching a 50 source/ load at
some common attenuation levels. The resistors corresponding to other attenuation levels may be
calculated from the equations.
dB = attenuation in decibels
Z = source/load impedance (resistive)
K>1
R3
VI
K =
= 10 dB/20
VO
2
R3 = Z K - 1
2K
R4 = Z K+ 1
K- 1

VI

R4

R4

attenuator

VO

Resistors for -section


Z=50.00
Attenuation
dB
K=Vi/Vo
R3
R4
1.0
1.12
5.77 869.55
2.0
1.26
11.61 436.21
3.0
1.41
17.61 292.40
4.0
1.58
23.85 220.97
6.0
2.00
37.35 150.48
10.0
3.16
71.15
96.25
20.0 10.00
247.50
61.11

Figure 1.14: Formulas for -section attenuator resistors, given K, the voltage attenuation ratio, and
ZI = ZO = 50 .
The above apply to the -attenuator below.
What resistor values would be required for both the attenuators for 10 dB of attenuation
matching a 50 source and load?

20

CHAPTER 1. AMPLIFIERS AND ACTIVE DEVICES

R3=71.2

Z
VI

=50

R4=
96.2 R4

VO

=50

attenuator
10 dB attenuator for matching input/output to Z= 50 .
Figure 1.15: 10 dB -section attenuator example for matching a 50 source and load.
The 10 dB corresponds to a voltage attenuation ratio of K=3.16 in the next to last line of the
above table. Transfer the resistor values in that line to the resistors on the schematic diagram in
Figure 1.15.

1.7.4

L-section attenuator

The table in Figure 1.16 lists resistor values for the L attenuators to match a 50 source/ load.
The table in Figure 1.17 lists resistor values for an alternate form. Note that the resistor values are
not the same.
dB = attenuation in decibels
Z = source/load impedance (resistive)
K>1
R

5
VI
= 10 dB/20
VO
VI
VO

R6
R5 = Z K- 1

K
R6 = Z
L attenuator
(K-1)

K =

Resistors for L-section


Z=,50.00
Attenuation L
dB
K=Vi/Vo
R5
R6
1.0
1.12
5.44 409.77
2.0
1.26 10.28 193.11
3.0
1.41 14.60 121.20
4.0
1.58 18.45
85.49
6.0
2.00 24.94
50.24
10.0
3.16 34.19
23.12
20.0
10.00 45.00
5.56

Figure 1.16: L-section attenuator table for 50 source and load impedance.
The above apply to the L attenuator below.

1.7.5

Bridged T attenuator

The table in Figure 1.18 lists resistor values for the bridged T attenuators to match a 50 source
and load. The bridged-T attenuator is not often used. Why not?

1.7.6

Cascaded sections

Attenuator sections can be cascaded as in Figure 1.19 for more attenuation than may be available
from a single section. For example two 10 db attenuators may be cascaded to provide 20 dB of

1.7.

ATTENUATORS

21

dB = attenuation in decibels
Z = source/load impedance (resistive)
K>1
R7
V
K = I = 10 dB/20
VO
VO
VI
R7 = Z(K-1)
R8

K
R8 = Z K- 1
L attenuator

Resistors for L-section


Z=50.00
Attenuation
dB K=Vi/Vo
R7
R8
1.0
1.12
6.10 459.77
2.0
1.26 12.95 243.11
3.0
1.41 20.63 171.20
4.0
1.58 29.24 135.49
6.0
2.00 49.76 100.24
10.0
3.16 108.11
73.12
20.0
10.00 450.00
55.56

Figure 1.17: Alternate form L-section attenuator table for 50 source and load impedance.

dB = attenuation in decibels
Z = source/load impedance (resistive)
K>1
VI
= 10 dB/20
VO
R6 = Z
(K-1)
R7 = Z(K-1)
K =

VI

R7

VO
R6

Bridged T attenuator

Resistors for bridged T


Z=50.00
Attenuation
dB K=Vi/Vo
R7
R6
1.0
1.12
6.10 409.77
2.0
1.26 12.95 193.11
3.0
1.41 20.63 121.20
4.0
1.58 29.24
85.49
6.0
2.00 49.76
50.24
10.0
3.16 108.11
23.12
20.0
10.00 450.00
5.56

Figure 1.18: Formulas and abbreviated table for bridged-T attenuator section, Z = 50 .

22

CHAPTER 1. AMPLIFIERS AND ACTIVE DEVICES

attenuation, the dB values being additive. The voltage attenuation ratio K or V I /VO for a 10 dB
attenuator section is 3.16. The voltage attenuation ratio for the two cascaded sections is the product
of the two Ks or 3.16x3.16=10 for the two cascaded sections.

section 1

section 2

Figure 1.19: Cascaded attenuator sections: dB attenuation is additive.


Variable attenuation can be provided in discrete steps by a switched attenuator. The example
Figure 1.20, shown in the 0 dB position, is capable of 0 through 7 dB of attenuation by additive
switching of none, one or more sections.
S1

4 dB

S2

2 dB

S3

1 dB

Figure 1.20: Switched attenuator: attenuation is variable in discrete steps.


The typical multi section attenuator has more sections than the above figure shows. The addition
of a 3 or 8 dB section above enables the unit to cover to 10 dB and beyond. Lower signal levels are
achieved by the addition of 10 dB and 20 dB sections, or a binary multiple 16 dB section.

1.7.7

RF attenuators

For radio frequency (RF) work (<1000 Mhz), the individual sections must be mounted in shielded
compartments to thwart capacitive coupling if lower signal levels are to be achieved at the highest
frequencies. The individual sections of the switched attenuators in the previous section are mounted
in shielded sections. Additional measures may be taken to extend the frequency range to beyond
1000 Mhz. This involves construction from special shaped lead-less resistive elements.
A coaxial T-section attenuator consisting of resistive rods and a resistive disk is shown in Figure 1.21. This construction is usable to a few gigahertz. The coaxial version would have one
resistive rod between two resistive disks in the coaxial line as in Figure 1.22.
RF connectors, not shown, are attached to the ends of the above T and attenuators. The
connectors allow individual attenuators to be cascaded, in addition to connecting between a source
and load. For example, a 10 dB attenuator may be placed between a troublesome signal source and
an expensive spectrum analyzer input. Even though we may not need the attenuation, the expensive
test equipment is protected from the source by attenuating any overvoltage.
Summary: Attenuators

1.7.

ATTENUATORS

23

metalic conductor

resistive disc
resistive rod
Coaxial T-attenuator for radio frequency work
Figure 1.21: Coaxial T-attenuator for radio frequency work.

metalic conductor

resistive rod
resistive disc
Coaxial -attenuator for radio frequency work
Figure 1.22: Coaxial -attenuator for radio frequency work.

24

CHAPTER 1. AMPLIFIERS AND ACTIVE DEVICES


An attenuator reduces an input signal to a lower level.
The amount of attenuation is specified in decibels (dB). Decibel values are additive for cascaded
attenuator sections.
dB from power ratio:

dB = 10 log10 (PI / PO )

dB from voltage ratio:

dB = 20 log10 (VI / VO )

T and section attenuators are the most common circuit configurations.

Contributors
Contributors to this chapter are listed in chronological order of their contributions, from most recent
to first. See Appendix 2 (Contributor List) for dates and contact information.
Colin Barnard (November 2003): Correction regarding Alexander Graham Bells country of
origin (Scotland, not the United States).

Chapter 2

SOLID-STATE DEVICE THEORY


Contents

2.1

2.1

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

25

2.2

Quantum physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

26

2.3

Valence and Crystal structure . . . . . . . . . . . . . . . . . . . . . . .

39

2.4

Band theory of solids . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

45

2.5

Electrons and holes . . . . . . . . . . . . . . . . . . . . . . . . . . . .

47

2.6

The P-N junction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

52

2.7

Junction diodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

55

2.8

Bipolar junction transistors . . . . . . . . . . . . . . . . . . . . . . . . .

57

2.9

Junction field-effect transistors . . . . . . . . . . . . . . . . . . . . . . .

62

2.10

Insulated-gate field-effect transistors (MOSFET) . . . . . . . . . . . .

67

2.11

Thyristors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

70

2.12

Semiconductor manufacturing techniques . . . . . . . . . . . . . . . .

72

2.13

Superconducting devices . . . . . . . . . . . . . . . . . . . . . . . . . . .

76

2.14

Quantum devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

79

2.15

Semiconductor devices in SPICE

. . . . . . . . . . . . . . . . . . . . .

87

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

90

Introduction

This chapter will cover the physics behind the operation of semiconductor devices and show how
these principles are applied in several different types of semiconductor devices. Subsequent chapters
will deal primarily with the practical aspects of these devices in circuits and omit theory as much
as possible.
25

26

2.2

CHAPTER 2. SOLID-STATE DEVICE THEORY

Quantum physics
I think it is safe to say that no one understands quantum mechanics.
Physicist Richard P. Feynman

To say that the invention of semiconductor devices was a revolution would not be an exaggeration.
Not only was this an impressive technological accomplishment, but it paved the way for developments that would indelibly alter modern society. Semiconductor devices made possible miniaturized
electronics, including computers, certain types of medical diagnostic and treatment equipment, and
popular telecommunication devices, to name a few applications of this technology.
But behind this revolution in technology stands an even greater revolution in general science: the
field of quantum physics. Without this leap in understanding the natural world, the development of
semiconductor devices (and more advanced electronic devices still under development) would never
have been possible. Quantum physics is an incredibly complicated realm of science. This chapter is
but a brief overview. When scientists of Feynmans caliber say that no one understands [it], you
can be sure it is a complex subject. Without a basic understanding of quantum physics, or at least
an understanding of the scientific discoveries that led to its formulation, though, it is impossible to
understand how and why semiconductor electronic devices function. Most introductory electronics
textbooks Ive read try to explain semiconductors in terms of classical physics, resulting in more
confusion than comprehension.
Many of us have seen diagrams of atoms that look something like Figure 2.1.
e

= electron

P = proton
N = neutron
e

N
P P
N P
N P
P
N P
N N

Figure 2.1: Rutherford atom: negative electrons orbit a small positive nucleus.
Tiny particles of matter called protons and neutrons make up the center of the atom; electrons
orbit like planets around a star. The nucleus carries a positive electrical charge, owing to the presence
of protons (the neutrons have no electrical charge whatsoever), while the atoms balancing negative
charge resides in the orbiting electrons. The negative electrons are attracted to the positive protons

2.2.

QUANTUM PHYSICS

27

a
pa lp
rt ha
ic
l

es

just as planets are gravitationally attracted by the Sun, yet the orbits are stable because of the
electrons motion. We owe this popular model of the atom to the work of Ernest Rutherford, who
around the year 1911 experimentally determined that atoms positive charges were concentrated in
a tiny, dense core rather than being spread evenly about the diameter as was proposed by an earlier
researcher, J.J. Thompson.
Rutherfords scattering experiment involved bombarding a thin gold foil with positively charged
alpha particles as in Figure 2.2. Young graduate students H. Geiger and E. Marsden experienced
unexpected results. A few Alpha particles were deflected at large angles. A few Alpha particles
were back-scattering, recoiling at nearly 180o . Most of the particles passed through the gold foil
undeflected, indicating that the foil was mostly empty space. The fact that a few alpha particles
experienced large deflections indicated the presence of a miniscule positively charged nucleus.

Gol

d fo

il

Figure 2.2: Rutherford scattering: a beam of alpha particles is scattered by a thin gold foil.
Although Rutherfords atomic model accounted for experimental data better than Thompsons,
it still wasnt perfect. Further attempts at defining atomic structure were undertaken, and these
efforts helped pave the way for the bizarre discoveries of quantum physics. Today our understanding
of the atom is quite a bit more complex. Nevertheless, despite the revolution of quantum physics
and its contribution to our understanding of atomic structure, Rutherfords solar-system picture of
the atom embedded itself in the popular conscience to such a degree that it persists in some areas
of study even when inappropriate.
Consider this short description of electrons in an atom, taken from a popular electronics textbook:
Orbiting negative electrons are therefore attracted toward the positive nucleus, which
leads us to the question of why the electrons do not fly into the atoms nucleus. The
answer is that the orbiting electrons remain in their stable orbit because of two equal
but opposite forces. The centrifugal outward force exerted on the electrons because of
the orbit counteracts the attractive inward force (centripetal) trying to pull the electrons
toward the nucleus because of the unlike charges.
In keeping with the Rutherford model, this author casts the electrons as solid chunks of matter
engaged in circular orbits, their inward attraction to the oppositely charged nucleus balanced by
their motion. The reference to centrifugal force is technically incorrect (even for orbiting planets),
but is easily forgiven because of its popular acceptance: in reality, there is no such thing as a force

28

CHAPTER 2. SOLID-STATE DEVICE THEORY

an
Lymies
ser

Bra
seri cket
es

Pasche
series n

Balmer
series

pushing any orbiting body away from its center of orbit. It seems that way because a bodys inertia
tends to keep it traveling in a straight line, and since an orbit is a constant deviation (acceleration)
from straight-line travel, there is constant inertial opposition to whatever force is attracting the
body toward the orbit center (centripetal), be it gravity, electrostatic attraction, or even the tension
of a mechanical link.
The real problem with this explanation, however, is the idea of electrons traveling in circular
orbits in the first place. It is a verifiable fact that accelerating electric charges emit electromagnetic
radiation, and this fact was known even in Rutherfords time. Since orbiting motion is a form of
acceleration (the orbiting object in constant acceleration away from normal, straight-line motion),
electrons in an orbiting state should be throwing off radiation like mud from a spinning tire. Electrons
accelerated around circular paths in particle accelerators called synchrotrons are known to do this,
and the result is called synchrotron radiation. If electrons were losing energy in this way, their orbits
would eventually decay, resulting in collisions with the positively charged nucleus. Nevertheless, this
doesnt ordinarily happen within atoms. Indeed, electron orbits are remarkably stable over a wide
range of conditions.
Furthermore, experiments with excited atoms demonstrated that electromagnetic energy emitted by an atom only occurs at certain, definite frequencies. Atoms that are excited by outside
influences such as light are known to absorb that energy and return it as electromagnetic waves
of specific frequencies, like a tuning fork that rings at a fixed pitch no matter how it is struck.
When the light emitted by an excited atom is divided into its constituent frequencies (colors) by a
prism, distinct lines of color appear in the spectrum, the pattern of spectral lines being unique to
that element. This phenomenon is commonly used to identify atomic elements, and even measure
the proportions of each element in a compound or chemical mixture. According to Rutherfords
solar-system atomic model (regarding electrons as chunks of matter free to orbit at any radius) and
the laws of classical physics, excited atoms should return energy over a virtually limitless range
of frequencies rather than a select few. In other words, if Rutherfords model were correct, there
would be no tuning fork effect, and the light spectrum emitted by any atom would appear as a
continuous band of colors rather than as a few distinct lines.

n=6

n=5
n=4

discharge lamp
slit

4102 A
4340
4861

H
H
H
H

n=3
n=2 (L)
n=1 (K)
nucleus

6563

Balmer series

Figure 2.3: Bohr hydrogen atom (with orbits drawn to scale) only allows electrons to inhabit discrete
orbitals. Electrons falling from n=3,4,5, or 6 to n=2 accounts for Balmer series of spectral lines.
A pioneering researcher by the name of Niels Bohr attempted to improve upon Rutherfords
model after studying in Rutherfords laboratory for several months in 1912. Trying to harmonize

2.2.

QUANTUM PHYSICS

29

the findings of other physicists (most notably, Max Planck and Albert Einstein), Bohr suggested
that each electron had a certain, specific amount of energy, and that their orbits were quantized such
that each may occupy certain places around the nucleus, as marbles fixed in circular tracks around
the nucleus rather than the free-ranging satellites each were formerly imagined to be. (Figure 2.3)
In deference to the laws of electromagnetics and accelerating charges, Bohr alluded to these orbits
as stationary states to escape the implication that they were in motion.
Although Bohrs ambitious attempt at re-framing the structure of the atom in terms that agreed
closer to experimental results was a milestone in physics, it was not complete. His mathematical
analyzes produced better predictions of experimental events than analyzes belonging to previous
models, but there were still some unanswered questions about why electrons should behave in such
strange ways. The assertion that electrons existed in stationary, quantized states around the nucleus
accounted for experimental data better than Rutherfords model, but he had no idea what would
force electrons to manifest those particular states. The answer to that question had to come from
another physicist, Louis de Broglie, about a decade later.
De Broglie proposed that electrons, as photons (particles of light) manifested both particlelike and wave-like properties. Building on this proposal, he suggested that an analysis of orbiting
electrons from a wave perspective rather than a particle perspective might make more sense of their
quantized nature. Indeed, another breakthrough in understanding was reached.

node

node

node

antinode

antinode

Figure 2.4: String vibrating at resonant frequency between two fixed points forms standing wave.
The atom according to de Broglie consisted of electrons existing as standing waves, a phenomenon
well known to physicists in a variety of forms. As the plucked string of a musical instrument
(Figure ??.
Electrons only could exist in certain, definite orbits around the nucleus because those were the
only distances where the wave ends would match. In any other radius, the wave should destructively
interfere with itself and thus cease to exist.
De Broglies hypothesis gave both mathematical support and a convenient physical analogy
to account for the quantized states of electrons within an atom, but his atomic model was still
incomplete. Within a few years, though, physicists Werner Heisenberg and Erwin Schrodinger,
working independently of each other, built upon de Broglies concept of a matter-wave duality to
create more mathematically rigorous models of subatomic particles.
This theoretical advance from de Broglies primitive standing wave model to Heisenbergs matrix and Schrodingers differential equation models was given the name quantum mechanics, and it
introduced a rather shocking characteristic to the world of subatomic particles: the trait of probability, or uncertainty. According to the new quantum theory, it was impossible to determine the
exact position and exact momentum of a particle at the same time. The popular explanations of

30

CHAPTER 2. SOLID-STATE DEVICE THEORY

ti

an

n
nti

tin

no

tin

od

od

node
node
nucleus

node

nucleus
node

node

od
tin
an

(a)

antinode

antinode

node

node

node

an

od

an

de

node
node

od

an

in
nt

de

tin

(b)

od

no
nti

Figure 2.5: Orbiting electron as standing wave around the nucleus, (a) two cycles per orbit, (b)
three cycles per orbit.
this uncertainty principle was that it was a measurement error usually caused by the process of
measurement (i.e. by attempting to precisely measure the position of an electron, you interfere
with its momentum and thus cannot know what it was before the position measurement was taken,
and vice versa). The startling implication of quantum mechanics is that particles do not actually
have precise positions and momenta, but rather balance the two quantities in a such way that their
combined uncertainties never diminish below a certain minimum value.
This form of uncertainty relationship exists in areas other than quantum mechanics. As discussed in the Mixed-Frequency AC Signals chapter in volume II of this book series, there is a
mutually exclusive relationship between the certainty of a waveforms time-domain data and its
frequency-domain data. In simple terms, the more precisely we know its constituent frequency(ies),
the less precisely we know its amplitude in time, and vice versa. To quote myself:
A waveform of infinite duration (infinite number of cycles) can be analyzed with
absolute precision, but the less cycles available to the computer for analysis, the less
precise the analysis. . . The fewer times that a wave cycles, the less certain its frequency
is. Taking this concept to its logical extreme, a short pulse a waveform that doesnt
even complete a cycle actually has no frequency, but rather acts as an infinite range of
frequencies. This principle is common to all wave-based phenomena, not just AC voltages
and currents.
In order to precisely determine the amplitude of a varying signal, we must sample it over a
very narrow span of time. However, doing this limits our view of the waves frequency. Conversely,
to determine a waves frequency with great precision, we must sample it over many cycles, which
means we lose view of its amplitude at any given moment. Thus, we cannot simultaneously know the
instantaneous amplitude and the overall frequency of any wave with unlimited precision. Stranger
yet, this uncertainty is much more than observer imprecision; it resides in the very nature of the wave.
It is not as though it would be possible, given the proper technology, to obtain precise measurements

2.2.

QUANTUM PHYSICS

31

of both instantaneous amplitude and frequency at once. Quite literally, a wave cannot have both a
precise, instantaneous amplitude, and a precise frequency at the same time.
The minimum uncertainty of a particles position and momentum expressed by Heisenberg and
Schrodinger has nothing to do with limitation in measurement; rather it is an intrinsic property of
the particles matter-wave dual nature. Electrons, therefore, do not really exist in their orbits as
precisely defined bits of matter, or even as precisely defined waveshapes, but rather as clouds
the technical term is wavefunction of probability distribution, as if each electron were spread or
smeared over a range of positions and momenta.
This radical view of electrons as imprecise clouds at first seems to contradict the original principle
of quantized electron states: that electrons exist in discrete, defined orbits around atomic nuclei.
It was, after all, this discovery that led to the formation of quantum theory to explain it. How
odd it seems that a theory developed to explain the discrete behavior of electrons ends up declaring
that electrons exist as clouds rather than as discrete pieces of matter. However, the quantized
behavior of electrons does not depend on electrons having definite position and momentum values,
but rather on other properties called quantum numbers. In essence, quantum mechanics dispenses
with commonly held notions of absolute position and absolute momentum, and replaces them with
absolute notions of a sort having no analogue in common experience.
Even though electrons are known to exist in ethereal, cloud-like forms of distributed probability
rather than as discrete chunks of matter, those clouds have other characteristics that are discrete.
Any electron in an atom can be described by four numerical measures (the previously mentioned
quantum numbers), called the Principal, Angular Momentum, Magnetic, and Spin numbers.
The following is a synopsis of each of these numbers meanings:
Principal Quantum Number: Symbolized by the letter n, this number describes the shell
that an electron resides in. An electron shell is a region of space around an atoms nucleus that
electrons are allowed to exist in, corresponding to the stable standing wave patterns of de Broglie
and Bohr. Electrons may leap from shell to shell, but cannot exist between the shell regions.
The principle quantum number must be a positive integer (a whole number, greater than or
equal to 1). In other words, principle quantum number for an electron cannot be 1/2 or -3. These
integer values were not arrived at arbitrarily, but rather through experimental evidence of light spectra: the differing frequencies (colors) of light emitted by excited hydrogen atoms follow a sequence
mathematically dependent on specific, integer values as illustrated in Figure 2.3.
Each shell has the capacity to hold multiple electrons. An analogy for electron shells is the
concentric rows of seats of an amphitheater. Just as a person seated in an amphitheater must
choose a row to sit in (one cannot sit between rows), electrons must choose a particular shell to
sit in. As in amphitheater rows, the outermost shells are hold more electrons than the inner shells.
Also, electrons tend to seek the lowest available shell, as people in an amphitheater seek the closest
seat to the center stage. The higher the shell number, the greater the energy of the electrons in it.
The maximum number of electrons that any shell may hold is described by the equation 2n 2 ,
where n is the principle quantum number. Thus, the first shell (n=1) can hold 2 electrons; the
second shell (n=2) 8 electrons, and the third shell (n=3) 18 electrons. (Figure 2.6)
Electron shells in an atom were formerly designated by letter rather than by number. The first
shell (n=1) was labeled K, the second shell (n=2) L, the third shell (n=3) M, the fourth shell (n=4)
N, the fifth shell (n=5) O, the sixth shell (n=6) P, and the seventh shell (n=7) Q.
Angular Momentum Quantum Number: A shell, is composed of subshells. One might be
inclined to think of subshells as simple subdivisions of shells, as lanes dividing a road. The subshells

32

CHAPTER 2. SOLID-STATE DEVICE THEORY


K

n= 1
2n2 = 2
observed fill = 2

2
8
8

3
18
18

4
32
32

18

18

Figure 2.6: Principal quantum number n and maximum number of electrons per shell both predicted
by 2(n2 ), and observed. Orbitals not to scale.
are much stranger. Subshells are regions of space where electron clouds are allowed to exist,
and different subshells actually have different shapes. The first subshell is shaped like a sphere,
(Figure 2.7(s) ) which makes sense when visualized as a cloud of electrons surrounding the atomic
nucleus in three dimensions. The second subshell, however, resembles a dumbbell, comprised of two
lobes joined together at a single point near the atoms center. (Figure 2.7(p) ) The third subshell
typically resembles a set of four lobes clustered around the atoms nucleus. These subshell shapes
are reminiscent of graphical depictions of radio antenna signal strength, with bulbous lobe-shaped
regions extending from the antenna in various directions. (Figure 2.7(d) )

y
x
z

x
1 of 3

1 of 1
px shown
py, pz similar

(s)

(p)

1 of 5
dx2 - y2 shown
dxy , dyz , dxz similar

(dx2 - y2)

1 of 5
dz2 shown

(dz2)

Figure 2.7: Orbitals: (s) Three fold symmetry. (p) Shown: sx , one of three possible orientations
(sx , sy , sz ), about their respective axes. (d) Shown: dx 2 -y x 2 similar to dxz , dyz , dxz . Shown: dz 2 .
Possible d-orbital orientations: five.
Valid angular momentum quantum numbers are positive integers like principal quantum numbers,
but also include zero. These quantum numbers for electrons are symbolized by the letter l. The
number of subshells in a shell is equal to the shells principal quantum number. Thus, the first shell
(n=1) has one subshell, numbered 0; the second shell (n=2) has two subshells, numbered 0 and 1;
the third shell (n=3) has three subshells, numbered 0, 1, and 2.
An older convention for subshell description used letters rather than numbers. In this notation,
the first subshell (l=0) was designated s, the second subshell (l=1) designated p, the third subshell

2.2.

QUANTUM PHYSICS

33

(l=2) designated d, and the fourth subshell (l=3) designated f. The letters come from the words
sharp, principal (not to be confused with the principal quantum number, n), diffuse, and fundamental. You will still see this notational convention in many periodic tables, used to designate the
electron configuration of the atoms outermost, or valence, shells. (Figure 2.8)
SL

SM
PL

PM

SN
PN
DM
DM SO

SK

18

18

10 2

10 1

electrons

n=

K
1

L
2

M
3
(a)

N
4

O
5

l=
n=
spectroscopic
notation

0
1

0, 1
2

0, 1, 2
3

0, 1, 2
4

0
5

1s2 2s22p6 3s23p63d10 4s24p64d10 5s1

(b)

Figure 2.8: (a) Bohr representation of Silver atom, (b) Subshell representation of Ag with division
of shells into subshells (angular quantum number l). This diagram implies nothing about the actual
position of electrons, but represents energy levels.
Magnetic Quantum Number: The magnetic quantum number for an electron classifies which
orientation its subshell shape is pointed. The lobes for subshells point in multiple directions. These
different orientations are called orbitals. For the first subshell (s; l=0), which resembles a sphere
pointing in no direction, so there is only one orbital. For the second (p; l=1) subshell in each shell,
which resembles dumbbells point in three possible directions. Think of three dumbbells intersecting
at the origin, each oriented along a different axis in a three-axis coordinate space.
Valid numerical values for this quantum number consist of integers ranging from -l to l, and
are symbolized as ml in atomic physics and lz in nuclear physics. To calculate the number of
orbitals in any given subshell, double the subshell number and add 1 (2l + 1). For example, the first
subshell (l=0) in any shell contains a single orbital, numbered 0; the second subshell (l=1) in any
shell contains three orbitals, numbered -1, 0, and 1; the third subshell (l=2) contains five orbitals,
numbered -2, -1, 0, 1, and 2; and so on.
Like principal quantum numbers, the magnetic quantum number arose directly from experimental
evidence: The Zeeman effect, the division of spectral lines by exposing an ionized gas to a magnetic
field, hence the name magnetic quantum number.
Spin Quantum Number: Like the magnetic quantum number, this property of atomic electrons was discovered through experimentation. Close observation of spectral lines revealed that each
line was actually a pair of very closely-spaced lines, and this so-called fine structure was hypothesized
to result from each electron spinning on an axis as if a planet. Electrons with different spins
would give off slightly different frequencies of light when excited. The name spin was assigned to
this quantum number. The concept of a spinning electron is now obsolete, being better suited to
the (incorrect) view of electrons as discrete chunks of matter rather than as clouds; but, the name
remains.

34

CHAPTER 2. SOLID-STATE DEVICE THEORY

Spin quantum numbers are symbolized as ms in atomic physics and sz in nuclear physics. For
each orbital in each subshell in each shell, there may be two electrons, one with a spin of +1/2 and
the other with a spin of -1/2.
The physicist Wolfgang Pauli developed a principle explaining the ordering of electrons in an
atom according to these quantum numbers. His principle, called the Pauli exclusion principle, states
that no two electrons in the same atom may occupy the exact same quantum states. That is, each
electron in an atom has a unique set of quantum numbers. This limits the number of electrons that
may occupy any given orbital, subshell, and shell.
Shown here is the electron arrangement for a hydrogen atom:

K shell
(n = 1)

subshell orbital
(l)
(ml)
0

spin
(ms)
1

/2

One electron

Hydrogen
Atomic number (Z) = 1
(one proton in nucleus)
Spectroscopic notation: 1s1
With one proton in the nucleus, it takes one electron to electrostatically balance the atom (the
protons positive electric charge exactly balanced by the electrons negative electric charge). This
one electron resides in the lowest shell (n=1), the first subshell (l=0), in the only orbital (spatial
orientation) of that subshell (ml =0), with a spin value of 1/2. A common method of describing
this organization is by listing the electrons according to their shells and subshells in a convention
called spectroscopic notation. In this notation, the shell number is shown as an integer, the subshell
as a letter (s,p,d,f), and the total number of electrons in the subshell (all orbitals, all spins) as a
superscript. Thus, hydrogen, with its lone electron residing in the base level, is described as 1s 1 .
Proceeding to the next atom (in order of atomic number), we have the element helium:

subshell orbital
(l)
(ml)
K shell
(n = 1)

0
0

spin
(ms)

-1/2

electron

electron

/2

Helium
Atomic number (Z) = 2
(two protons in nucleus)
Spectroscopic notation: 1s2
A helium atom has two protons in the nucleus, and this necessitates two electrons to balance
the double-positive electric charge. Since two electrons one with spin=1/2 and the other with
spin=-1/2 fit into one orbital, the electron configuration of helium requires no additional subshells

2.2.

QUANTUM PHYSICS

35

or shells to hold the second electron.

However, an atom requiring three or more electrons will require additional subshells to hold all
electrons, since only two electrons will fit into the lowest shell (n=1). Consider the next atom in the
sequence of increasing atomic numbers, lithium:

subshell orbital
(l)
(ml)

spin
(ms)

L shell
(n = 2)

/2

electron

K shell
(n = 1)

-1/2

electron

electron

/2

Lithium
Atomic number (Z) = 3
Spectroscopic notation: 1s22s1

An atom of lithium uses a fraction of the L shells (n=2) capacity. This shell actually has a total
capacity of eight electrons (maximum shell capacity = 2n2 electrons). If we examine the organization
of the atom with a completely filled L shell, we will see how all combinations of subshells, orbitals,
and spins are occupied by electrons:

36

CHAPTER 2. SOLID-STATE DEVICE THEORY

subshell orbital
(l)
(ml)
1

-1/2

- /2

-1

- /2

-1

- /2

-1/2

1
L shell
(n = 2)

K shell
(n = 1)

spin
(ms)

/2

/2

p subshell
(l = 1)
6 electrons

/2
/2
/2

s subshell
(l = 0)
2 electrons
s subshell
(l = 0)
2 electrons

Neon
Atomic number (Z) = 10
Spectroscopic notation: 1s22s22p6
Often, when the spectroscopic notation is given for an atom, any shells that are completely
filled are omitted, and the unfilled, or the highest-level filled shell, is denoted. For example, the
element neon (shown in the previous illustration), which has two completely filled shells, may be
spectroscopically described simply as 2p6 rather than 1s2 2s2 2p6 . Lithium, with its K shell completely
filled and a solitary electron in the L shell, may be described simply as 2s 1 rather than 1s2 2s1 .
The omission of completely filled, lower-level shells is not just a notational convenience. It
also illustrates a basic principle of chemistry: that the chemical behavior of an element is primarily
determined by its unfilled shells. Both hydrogen and lithium have a single electron in their outermost
shells (1s1 and 2s1 , respectively), and this gives the two elements some similar properties. Both are
highly reactive, and reactive in much the same way (bonding to similar elements in similar modes).
It matters little that lithium has a completely filled K shell underneath its almost-vacant L shell:
the unfilled L shell is the shell that determines its chemical behavior.
Elements having completely filled outer shells are classified as noble, and are distinguished by
almost complete non-reactivity with other elements. These elements used to be classified as inert,
when it was thought that these were completely unreactive, but are now known to form compounds
with other elements under specific conditions.
Since elements with identical electron configurations in their outermost shell(s) exhibit similar
chemical properties, Dimitri Mendeleev organized the different elements in a table accordingly. Such
a table is known as a periodic table of the elements, and modern tables follow this general form in
Figure 2.9.

2.2.

13

IA

H
1
Hydrogen
1.00794
1s1
Li

Periodic Table of the Elements

Figure 2.9: Periodic table of chemical elements.

IIA

Group new

3 Be
4
Lithium
Beryllium
6.941
9.012182

Symbol

2s1

K
19
Potassium
39.0983

Name

2s2

Electron
configuration

Na
11 Mg
12
Sodium Magnesium
22.989768 24.3050

IA

4s1

Metalloids
Nonmetals
IVA 15
VA 16
VIA 17 VIIA
13
IIIA 14
B
5 C
6 N
7 O
8 F
9
Boron
Carbon
Nitrogen
Oxygen
Fluorine
10.81
12.011
14.0067
15.9994
18.9984

Group old
Atomic number
Atomic mass
(averaged according to
occurence on earth)

2p1

2p2

2p3

2p4

2p5

VIIIA

He
2
Helium
4.00260
1s2
Ne
10
Neon
20.179
2p6

Al
13 Si
14 P
15 S
16 Cl
17 Ar
18
Aluminum
Silicon Phosphorus Sulfur
Chlorine
Argon
26.9815
28.0855
30.9738
32.06
35.453
39.948

QUANTUM PHYSICS

Metals
IB 12
3s1
3s2
3
IIIB 4
IVB 5
VB 6
VIB 7
VIIB 8
IIB
3p1
3p2
3p3
3p4
3p5
3p6
VIIIB 9 VIIIB 10 VIIIB 11
21 Ti
22 V
23 Cr
24 Mn
25 Fe
26 Co
27 Ni
28 Cu
29 Zn
30 Ga
31 Ge
32 As
33 Se
34 Br
35 Kr
36
K
19 Ca
20 Sc
Scandium Titanium
Vanadium Chromium Manganese
Iron
Cobalt
Nickel
Copper
Zinc
Gallium Germanium Arsenic
Selenium Bromine
Krypton
Potassium Calcium
50.9415
55.847
63.546
74.92159
78.96
79.904
83.80
39.0983
40.078
44.955910
47.88
51.9961
54.93805
58.93320
58.69
65.39
69.723
72.61
4s1

4s2

3d14s2

3d24s2

3d34s2

3d54s1

3d54s2

3d64s2

3d74s2

3d84s2

3d104s1

3d104s2

4p1

4p2

4p3

4p4

4p5

4p6

Rb
38 Y
40 Nb
41 Mo
42 Tc
43 Ru
44 Rh
45 Pd
46 Ag
47 Cd
48 In
49 Sn
50 Sb
51 Te
52 I
53 Xe
54
37 Sr
39 Zr
Zirconium Niobium Molybdenum Technetium Ruthenium Rhodium
Palladium
Silver
Cadmium
Indium
Tin
Antimony Tellurium
Iodine
Xenon
Rubidium Strontium
Yttrium
87.62
91.224
92.90638
(98)
101.07
106.42
107.8682
112.411
114.82
118.710
121.75
127.60
126.905
131.30
85.4678
88.90585
95.94
102.90550
5s1

5s2

4d15s2

4d25s2

4d45s1

4d55s1

4d55s2

4d75s1

4d85s1

4d105s0

4d105s1

4d105s2

5p1

5p2

5p3

5p4

5p5

5p6

Cs
55 Ba
56
57 - 71
Hf
72 Ta
73 W
74 Re
75 Os
76 Ir
77 Pt
78 Au
79 Hg
80 Tl
81 Pb
82 Bi
83 Po
84 At
85 Rn
86
Cesium
Barium
Lanthanide Hafnium
Tantalum Tungsten Rhenium
Osmium
Iridium
Platinum
Gold
Mercury
Thallium
Lead
Bismuth
Polonium
Astatine
Radon
series
132.90543
137.327
178.49
180.9479
183.85
186.207
190.2
192.22
195.08
196.96654
200.59
204.3833
207.2
208.98037
(209)
(210)
(222)
6s1

6s2

Fr
87 Ra
88
Francium
Radium
(223)
(226)
1

7s

5d26s2
89 - 103
Actinide
series

Unq

(261)
6d 7s

Lanthanide
series

6d 7s

Unnilhexium

(263)
4

6d 7s

5d56s2

106 Uns

5d66s2

107

5d76s2

108

5d96s1

5d106s1

5d106s2

6p1

6p2

6p3

6p4

6p5

6p6

109

Unnilseptium

(262)

La
57 Ce
58 Pr
59 Nd
60 Pm
61 Sm
62 Eu
63 Gd
64 Tb
65 Dy
66 Ho
67 Er
68 Tm
69 Yb
70 Lu
71
Lanthanum Cerium Praseodymium Neodymium Promethium Samarium Europium Gadolinium Terbium Dysprosium Holmium
Erbium
Thulium
Ytterbium
Lutetium
138.9055
140.115 140.90765
144.24
(145)
150.36
151.965
157.25
158.92534
162.50
164.93032
167.26
168.93421
173.04
174.967
5d16s2

Actinide
series

(262)
3

5d46s2

105 Unh

Unnilquadium Unnilpentium
2

7s

5d36s2

104 Unp

4f15d16s2

4f36s2

4f46s2

4f56s2

4f66s2

4f76s2

4f75d16s2

4f96s2

4f106s2

4f116s2

4f126s2

4f136s2

4f146s2

4f145d16s2

Ac
89 Th
90 Pa
91 U
92 Np
93 Pu
94 Am
95 Cm
96 Bk
97 Cf
98 Es
99 Fm
100 Md
101 No
102 Lr
103
Actinium
Thorium Protactinium Uranium Neptunium Plutonium Americium
Curium
Berkelium Californium Einsteinium Fermium Mendelevium Nobelium Lawrencium
(227)
232.0381 231.03588 238.0289
(237)
(244)
(243)
(247)
(247)
(251)
(252)
(257)
(258)
(259)
(260)
6d17s2

6d27s2

5f26d17s2

5f36d17s2

5f46d17s2

5f66d07s2

5f76d07s2

5f76d17s2

5f96d07s2

5f106d07s2

5f116d07s2

5f126d07s2

5f136d07s2

6d07s2

6d17s2

37

38

CHAPTER 2. SOLID-STATE DEVICE THEORY

Dmitri Mendeleev, a Russian chemist, was the first to develop a periodic table of the elements.
Although Mendeleev organized his table according to atomic mass rather than atomic number, and
produced a table that was not quite as useful as modern periodic tables, his development stands as an
excellent example of scientific proof. Seeing the patterns of periodicity (similar chemical properties
according to atomic mass), Mendeleev hypothesized that all elements should fit into this ordered
scheme. When he discovered empty spots in the table, he followed the logic of the existing order
and hypothesized the existence of heretofore undiscovered elements. The subsequent discovery of
those elements granted scientific legitimacy to Mendeleevs hypothesis, further discoveries leading
to the form of the periodic table we use today.
This is how science should work: hypotheses followed to their logical conclusions, and accepted,
modified, or rejected as determined by the agreement of experimental data to those conclusions.
Any fool may formulate a hypothesis after-the-fact to explain existing experimental data, and many
do. What sets a scientific hypothesis apart from post hoc speculation is the prediction of future
experimental data yet uncollected, and the possibility of disproof as a result of that data. To boldly
follow a hypothesis to its logical conclusion(s) and dare to predict the results of future experiments
is not a dogmatic leap of faith, but rather a public test of that hypothesis, open to challenge from
anyone able to produce contradictory data. In other words, scientific hypotheses are alwaysrisky
due to the claim to predict the results of experiments not yet conducted, and are therefore susceptible
to disproof if the experiments do not turn out as predicted. Thus, if a hypothesis successfully predicts
the results of repeated experiments, its falsehood is disproven.
Quantum mechanics, first as a hypothesis and later as a theory, has proven to be extremely
successful in predicting experimental results, hence the high degree of scientific confidence placed in
it. Many scientists have reason to believe that it is an incomplete theory, though, as its predictions
hold true more at micro physical scales than at macroscopic dimensions, but nevertheless it is a
tremendously useful theory in explaining and predicting the interactions of particles and atoms.
As you have already seen in this chapter, quantum physics is essential in describing and predicting many different phenomena. In the next section, we will see its significance in the electrical
conductivity of solid substances, including semiconductors. Simply put, nothing in chemistry or
solid-state physics makes sense within the popular theoretical framework of electrons existing as
discrete chunks of matter, whirling around atomic nuclei like miniature satellites. It is when electrons are viewed as wavefunctions existing in definite, discrete states that the regular and periodic
behavior of matter can be explained.
REVIEW:
Electrons in atoms exist in clouds of distributed probability, not as discrete chunks of matter
orbiting the nucleus like tiny satellites, as common illustrations of atoms show.
Individual electrons around an atomic nucleus seek unique states, described by four quantum numbers: the Principal Quantum Number, known as the shell ; the Angular Momentum
Quantum Number, known as the subshell ; the Magnetic Quantum Number, describing the orbital (subshell orientation); and the Spin Quantum Number, or simply spin. These states are
quantized, meaning that no in-between conditions exist for an electron other than those
states that fit into the quantum numbering scheme.
The Principal Quantum Number (n) describes the basic level or shell that an electron resides in.
The larger this number, the greater radius the electron cloud has from the atoms nucleus, and

2.3.

VALENCE AND CRYSTAL STRUCTURE

39

the greater that electrons energy. Principal quantum numbers are whole numbers (positive
integers).
The Angular Momentum Quantum Number (l ) describes the shape of the electron cloud within
a particular shell or level, and is often known as the subshell. There are as many subshells
(electron cloud shapes) in any given shell as that shells principal quantum number. Angular
momentum quantum numbers are positive integers beginning at zero and ending at one less
than the principal quantum number (n-1).
The Magnetic Quantum Number (ml ) describes which orientation a subshell (electron cloud
shape) has. Subshells may assume as many different orientations as the subshell number (l )
plus 1, and each unique orientation is called an orbital. These numbers are integers ranging
from the negative value of the subshell number (l ) through 0 to the positive value of the
subshell number.
The Spin Quantum Number (ms ) describes another property of an electron, and may be a
value of +1/2 or -1/2.
Paulis Exclusion Principle says that no two electrons in an atom may share the exact same
set of quantum numbers. Therefore, no more than two electrons may occupy each orbital
(spin=1/2 and spin=-1/2), 2l+1 orbitals in every subshell, and n subshells in every shell, and
no more.
Spectroscopic notation is a convention for denoting the electron configuration of an atom.
Shells are shown as whole numbers, followed by subshell letters (s,p,d,f), with superscripted
numbers totaling the number of electrons residing in each respective subshell.
An atoms chemical behavior is solely determined by the electrons in the unfilled shells. Lowlevel shells that are completely filled have little or no effect on the chemical bonding characteristics of elements.
Elements with completely filled electron shells are almost entirely unreactive, and are called
noble (formerly known as inert).

2.3

Valence and Crystal structure

Valence: The electrons in the outer most shell, or valence shell, are known as valence electrons.
These valence electrons are responsible for the chemical properties of the chemical elements. It
is these electrons which participate in chemical reactions with other elements. An over simplified
chemistry rule applicable to simple reactions is that atoms try to form a complete outer shell of
8 electrons (two for the L shell). Atoms may give away a few electrons to expose an underlying
complete shell. Atoms may accept a few electrons to complete the shell. These two processes form
ions from atoms. Atoms may even share electrons among atoms in an attempt to complete the outer
shell. This process forms molecular bonds. That is, atoms associate to form a molecule.
For example group I elements: Li, Na, K, Cu, Ag, and Au have a single valence electron.
(Figure 2.10) These elements all have similar chemical properties. These atoms readily give away
one electron to react with other elements. The ability to easily give away an electron makes these
elements excellent conductors.

40

CHAPTER 2. SOLID-STATE DEVICE THEORY

Li

Na

Cu

Ag

Au

Figure 2.10: Periodic table group IA elements: Li, Na, and K, and group IB elements: Cu, Ag, and
Au have one electron in the outer, or valence, shell, which is readily donated. Inner shell electrons:
For n= 1, 2, 3, 4; 2n2 = 2, 8, 18, 32.
Group VIIA elements: Fl, Cl, Br, and I all have 7 electrons in the outer shell. These elements
readily accept an electron to fill up the outer shell with a full 8 electrons. (Figure 2.11) If these
elements do accept an electron, a negative ion is formed from the neutral atom. These elements
which do not give up electrons are insulators.

Cl

Br

Figure 2.11: Periodic table group VIIA elements: F, Cl, Br, and I with 7 valence electrons readily
accept an electron in reactions with other elements.
For example, a Cl atom accepts an electron from an Na atom to become a Cl ion as shown
in Figure 2.12. An ion is a charged particle formed from an atom by either donating or accepting
an electron. As the Na atom donates an electron, it becomes a Na+ ion. This is how Na and Cl
atoms combine to form NaCl, table salt, which is actually Na+ Cl , a pair of ions. The Na+ and
Cl carrying opposite charges, attract one other.
+

=
Na

Cl

Na+

Cl-

Figure 2.12: Neutral Sodium atom donates an electron to neutral Chlorine atom forming Na + and
Cl ions.

2.3.

VALENCE AND CRYSTAL STRUCTURE

41

Sodium chloride crystallizes in the cubic structure shown in Figure 2.16. This model is not to
scale to show the three dimensional structure. The Na+ Cl ions are actually packed similar to
layers of stacked marbles. The easily drawn cubic crystal structure illustrates that a solid crystal
may contain charged particles.
Group VIIIA elements: He, Ne, Ar, Kr, Xe all have 8 electrons in the valence shell. (Figure 2.13)
That is, the valence shell is complete. These elements neither donate nor accept electrons. Nor do
they readily participate in chemical reactions. Group VIIIA elements do not readily combining with
other elements. Though chemist in recent years have forced Xe and Kr to form a few compounds,
this is not applicable to our discussion. These elements are good electrical insulators and are gases
at room temperature.

He

Ne

Ar

Kr

Xe

Figure 2.13: Group VIIIA elements: He, Ne, Ar, Kr, Xe are largely unreactive since the valence
shell is complete..
Group IVA elements: C, Si, Ge, having 4 electrons in the valence shell as shown in Figure 2.14
form compounds by sharing electrons with other elements without forming ions. This shared electron
bonding is known as covalent bonding. Note that the center atom (and the others by extension) has
completed its valence shell by sharing electrons. Note that the figure is a 2-d representation of
bonding, which is actually 3-d. It is this group, IVA, that we are interested in for its semiconducting
properties.

Si
(a)

Ge
(b)

Figure 2.14: (a) Group IVA elements: C, Si, Ge having 4 electrons in the valence shell, (b) complete
the valence shell by sharing electrons with other elements.
Crystal structure: Most inorganic substances form their atoms (or ions) into an ordered array
known as a crystal. The outer electron clouds of atoms interact in an orderly manner. Even metals

42

CHAPTER 2. SOLID-STATE DEVICE THEORY

are composed of crystals at the microscopic level. If a metal sample is given an optical polish,
then acid etched, the microscopic microcrystalline structure shows as in Figure 2.15. It is also
possible to purchase, at considerable expense, metallic single crystal specimens from specialized
suppliers. Polishing and etching such a specimen discloses no microcrystalline structure. Practically
all industrial metals are polycrystalline. Most modern semiconductors, on the other hand, are single
crystal devices. We are primarily interested in monocrystalline structures.

(a)

(b)

(c)

Figure 2.15: (a) Metal sample, (b) polished, (c) acid etched to show microcrystalline structure.
Many metals are soft and easily deformed by the various metal working techniques. The microcrystals are deformed in metal working. Also, the valence electrons are free to move about the
crystal lattice, and from crystal to crystal. The valence electrons do not belong to any particular
atom, but to all atoms.
The rigid crystal structure in Figure 2.16 is composed of a regular repeating pattern of positive Na
ions and negative Cl ions. Once the Na and Cl atoms have formed Na+ and Cl ions by transferring
an electron from Na to Cl, with no free electrons. Electrons are not free to move about the crystal
lattice, a difference compared with a metal. Nor are the ions free. Ions are fixed in place within the
crystal structure. Though, the ions are free to move about if the NaCl crystal is dissolved in water.
However, the crystal no longer exists. The regular, repeating structure is gone. Evaporation of the
water deposits the Na+ and Cl ions in the form of new crystals as the oppositely charged ions
attract each other. Ionic materials form crystal structures due to the strong electrostatic attraction
of the oppositely charged ions.
Semiconductors in Group IV also form crystals because of the tetrahedral bonding pattern of the
s2 p2 electrons about the atom, sharing electron-pair bonds to four adjacent atoms. (Figure 2.18(a) )
More correctly the four outer electrons: two in the s-orbital, (sz ) offset along the z-axis, and two in
the p-orbital (px and py ) hybridize to form four sp3 molecular orbitals. These four electron clouds
repell one another to equidistant tetrahedral spacing about the Si atom, attracted by the positive
nucleus as shown in Figure 2.17.
Every semiconductor atom, Si, Ge, or C (diamond) is chemically bonded to four other atoms by
covalent bonds, shared electron bonds. Two electrons may share an orbital if each have opposite spin
quantum numbers. Thus, an unpaired electron may share an orbital with an electron from another
atom. This corresponds to overlapping Figure 2.18(a) of the electron clouds, or bonding. Figure 2.18
(b) is one fourth of the volume of the diamond crystal structure unit cell shown in Figure 2.19 at
the origin. The bonds are particularly strong in diamond, decreasing in strength going down group
IV to silicon, and germanium. Silicon and germanium both form crystals with a diamond structure.
The diamond unit cell, the basic crystal building block, in Figure 2.19 shows four atoms (dark)
bonded to four others within the volume of the cell. This is equivalent to placing one of Figure 2.18(b)
at the origin in Figure 2.19, then placing three more on adjacent faces to fill the full cube. Six atoms

2.3.

VALENCE AND CRYSTAL STRUCTURE

43

Na+
Cl -

Figure 2.16: NaCl crystal having a cubic structure.

x
z
sz 2

px

py

sp3

Figure 2.17: Two s-orbital (sz ) electrons and two p-orbital (sx and sy ) electrons hybridize, (c)
forming four sp3 molecular orbitals.

(a)

(b)

Figure 2.18: (a) Tetrahedral bonding of Si atom. (b) leads to 1/4 of the cubic unit cell

44

CHAPTER 2. SOLID-STATE DEVICE THEORY

fall on the middle of each of the six cube faces, showing two bonds. The other two bonds to adjacent
cubes were omitted for clarity. Out of eight cube corners, four atoms bond to an atom within the
cube. Where are the other four atoms bonded? The other four bond to adjacent cubes of the crystal.
Keep in mind that even though four corner atoms show no bonds in the cube, all atoms within the
crystal are bonded in one giant molecule. A semiconductor crystal is built up from copies of this
unit cell.

Face centered atoms

Atom bonded to 4 others


Other atoms bonded to
chain in cube
Atoms bonded outside of
cube

Figure 2.19: Si, Ge, and C (diamond) form interleaved face centered cube.
The crystal is effectively one molecule. An atom covalent bonds to four others, which in turn bond
to four others, and so on. The crystal lattice is relatively stiff resisting deformation. Few electrons
free themselves for conduction about the crystal. A property of semiconductors is that once an
electron is freed, a positively charged empty space develops which also contributes to conduction.
REVIEW
Atoms try to form a complete outer, valence, shell of 8-electrons (2-electrons for the innermost
shell). Atoms may donate a few to expose an userlying shell of 8, accept a few to complete a
shell, or share electrons to complete a shell.
Atoms often form ordered arrays of ions or atoms in a rigid structure known as a crystal.
A neutral atom may form a positive ion by donating an electron.
A neutral atom may form a negative ion by accepting an electron
The group IVA semiconductors: C, Si, Ge crystallize into a diamond structure. Each atom in
the crystal is part of a giant molecule, bonding to four other atoms.

2.4.

BAND THEORY OF SOLIDS

45

Most semiconductor devices are manufactured from single crystals.

2.4

Band theory of solids

Quantum physics describes the states of electrons in an atom according to the four-fold scheme of
quantum numbers. The quantum numbers describe the allowable states electrons may assume in an
atom. To use the analogy of an amphitheater, quantum numbers describe how many rows and seats
are available. Individual electrons may be described by the combination of quantum numbers, like
a spectator in an amphitheater assigned to a particular row and seat.
Like spectators in an amphitheater moving between seats and rows, electrons may change their
statuses, given the presence of available spaces for them to fit, and available energy. Since shell
level is closely related to the amount of energy that an electron possesses, leaps between shell
(and even subshell) levels requires transfers of energy. If an electron is to move into a higher-order
shell, it requires that additional energy be given to the electron from an external source. Using
the amphitheater analogy, it takes an increase in energy for a person to move into a higher row of
seats, because that person must climb to a greater height against the force of gravity. Conversely,
an electron leaping into a lower shell gives up some of its energy, like a person jumping down into
a lower row of seats, the expended energy manifesting as heat and sound.
Not all leaps are equal. Leaps between different shells requires a substantial exchange of energy,
but leaps between subshells or between orbitals require lesser exchanges.
When atoms combine to form substances, the outermost shells, subshells, and orbitals merge,
providing a greater number of available energy levels for electrons to assume. When large numbers
of atoms are close to each other, these available energy levels form a nearly continuous band wherein
electrons may move as illustrated in Figure 2.20

Significant leap required


for an electron to move
to the next higher level
3p

Shorter leap
required

Overlap permits
electrons to freely
drift between bands

3p

3p
Overlap

3s
Single atom

3s

3s
Five atoms
in close proximity

Multitudes of atoms
in close proximity

Figure 2.20: Electron band overlap in metallic elements.


It is the width of these bands and their proximity to existing electrons that determines how
mobile those electrons will be when exposed to an electric field. In metallic substances, empty
bands overlap with bands containing electrons, meaning that electrons of a single atom may move
to what would normally be a higher-level state with little or no additional energy imparted. Thus,

46

CHAPTER 2. SOLID-STATE DEVICE THEORY

the outer electrons are said to be free, and ready to move at the beckoning of an electric field.
Band overlap will not occur in all substances, no matter how many atoms are in close to each
other. In some substances, a substantial gap remains between the highest band containing electrons
(the so-called valence band ) and the next band, which is empty (the so-called conduction band ). See
Figure 2.21. As a result, valence electrons are bound to their constituent atoms and cannot become
mobile within the substance without a significant amount of imparted energy. These substances are
electrical insulators.

Conduction band
Significant leap required
for an electron to enter
the conduction band and
travel through the material

"Energy gap"
Valence band

Multitudes of atoms
in close proximity
Figure 2.21: Electron band separation in insulating substances.
Materials that fall within the category of semiconductors have a narrow gap between the valence
and conduction bands. Thus, the amount of energy required to motivate a valence electron into the
conduction band where it becomes mobile is quite modest. (Figure 2.22)
semiconducting substance
metalic substance for reference

Conduction band
Small leap required
for an electron to enter
the conduction band and
travel through the material

Insignificant leap
for electron to
enter conduction
band

"Energy gap"
Valence band

(a)

(b)

Figure 2.22: Electron band separation in semiconducting substances, (a) multitudes of semiconducting close atoms still results in a significant band gap, (b) multitudes of close metal atoms for
reference.
At low temperatures, little thermal energy is available to push valence electrons across this gap,

2.5.

ELECTRONS AND HOLES

47

and the semiconducting material acts more as an insulator. at higher temperatures, though, the
ambient thermal energy becomes enough to force electrons across the gap, and the material will
increase conduction of electricity.
It is difficult to predict the conductive properties of a substance by examining the electron configurations of its constituent atoms. Although the best metallic conductors of electricity (silver, copper,
and gold) all have outer s subshells with a single electron, the relationship between conductivity
and valence electron count is not necessarily consistent:
Element

Specific resistance Electron


() at 20o Celsius configuration

Element

Specific resistance Electron


() at 20o Celsius configuration

Molybdenum (Mo) 32.12 cmil/ft

Silver (Ag)

9.546 cmil/ft

4d105s1

Copper (Cu)

10.09 cmil/ft

3d104s1

Zinc (Zn)

35.49 cmil/ft

3d104s2

Gold (Au)

13.32 cmil/ft

5d106s1

Nickel (Ni)

41.69 cmil/ft

3d84s2

Aluminum (Al)

15.94 cmil/ft

3p1

Iron (Fe)

57.81 cmil/ft

3d64s2

Tungsten (W)

31.76 cmil/ft

5d46s2

Platinum (Pt)

63.16 cmil/ft

5d96s1

4d55s1

The electron band configurations produced by compounds of different elements defies easy association with the electron configurations of its constituent elements.
REVIEW:
Energy is required to remove an electron from the valence band to a higher unoccupied band,
a conduction band. More energy is required to move between shells, less between subshells.
Since the valence and conduction bands overlap in metals, little energy removes an electron.
Metals are excellent conductors.
The large gap between the valence and conduction bands of an insulator requires high energy
to remove an electron. Thus, insulators do not conduct.
Semiconductors have a small non-overlapping gap between the valence and conduction bands.
Pure semiconductors are neither good insulators nor conductors. Semiconductors are semiconductive.

2.5

Electrons and holes

Pure semiconductors are relatively good insulators as compared with metals, though not nearly
as good as a true insulator like glass. To be usefull in semiconductor applications, the intrinsic
semiconductor, pure undoped semiconductor must have no more than one impurity atom in 10
billion semiconductor atoms. This is analogous to a grain of salt impurity in a railroad boxcar of
sugar. Impure, or dirty semiconductors are considerably more conductive, though not as good as
metals. Why might this be? To answer that question, we must look at the electron structure of
such materials in Figure 2.23.
Figure 2.23 (a) shows four electrons in the valence shell of a semiconductor forming covalent
bonds to four other atoms. This is a flattened, easier to draw, version of Figure 2.19. All electrons
of an atom are tied up in four covalent bonds, pairs of shared electrons. Electrons are not free to

48

CHAPTER 2. SOLID-STATE DEVICE THEORY

move about the crystal lattice. Thus, intrinsic, pure, semiconductors are relatively good insulators
as compared to metals.

(a)

hole

(b)

electron

Figure 2.23: (a) Intrinsic semiconductor is an insulator having a complete electron shell. (b) However,
thermal energy can create few electron hole pairs resulting in weak conduction.
Thermal energy may occasionally free an electron from the crystal lattice as in Figure 2.23 (b).
This electron is free for conduction about the crystal lattice. When the electron was freed, it left an
empty spot with a positive charge in the crystal lattice known as a hole. This hole is not fixed to the
lattice; but, is free to move about. The free electron and hole both contribute to conduction about
the crystal lattice. That is, the electron is free until it falls into a hole. This is called recombination.
If an external electric field is applied to the semiconductor, the electrons and holes will conduct
in opposite directions. Increasing temperature will increase the number of electrons and holes,
decreasing the resistance. This is opposite of metals, where resistance increases with temperature
by increasing the collisions of electrons with the crystal lattice. The number of electrons and holes in
an intrinsic semiconductor are equal. However, both carriers do not necessarily move with the same
velocity with the application of an external field. Another way of stating this is that the mobility is
not the same for electrons and holes.
Pure semiconductors, by themselves, are not particularly usefully. Though, semiconductors must
be refined to a high level of purity as a starting point prior the addition of specific impurities.
Semiconductor material pure to 1 part in 10 billion, may have specific impurities added at approximately 1 part per 10 million to increase the number of carriers. The addition of a desired impurity
to a semiconductor is known as doping. Doping increases the conductivity of a semiconductor so
that it is more comparable to a metal than an insulator.
It is possible to increase the number of negative charge carriers within the semiconductor crystal
lattice by doping with an electron donor like Phosphorus. Electron donors, also known as N-type
dopants include elements from group VA of the periodic table: nitrogen, phosphorus, arsenic, and
antimony. Nitrogen and phosphorus are N-type dopants for diamond. Phosphorus, arsenic, and
antimony are used with silicon.
The crystal lattice in Figure 2.24 (a) contains atoms having four electrons in the outer shell,
forming four covalent bonds to adjacent atoms. This is the anticipated crystal lattice. The addition
of a phosphorus atom with five electrons in the outer shell introduces an extra electron into the
lattice as compared with the silicon atom. The pentavalent impurity forms four covalent bonds to
four silicon atoms with four of the five electrons, fitting into the lattice with one electron left over.
Note that this spare electron is not strongly bonded to the lattice as the electrons of normal Si
atoms are. It is free to move about the crystal lattice, not being bound to the Phosphorus lattice

2.5.

ELECTRONS AND HOLES

49

site. Since we have doped at one part phosphorus in 10 million silicon atoms, few free electrons
were created compared with the numerous silicon atoms. However, many electrons were created
compared with the fewer electron-hole pairs in intrinsic silicon. Application of an external electric
field produces strong conduction in the doped semiconductor in the conduction band (above the
valence band). A heavier doping level produces stronger conduction. Thus, a poorly conducting
intrinsic semiconductor has been converted into a good electrical conductor.

B
Si

(a)

(b)

Si

Si

Si

Si

Si

Si

Si

Si

Si

Si

Si

Si

Si

Si

Si

Si

Si

Si
electron

Si
Si
Si
Si
(c)
hole
electron
hole movement
movement

Figure 2.24: (a) Outer shell electron configuration of donor N-type Phosphorus, Silicon (for reference), and acceptor P-type Boron. (b) N-type donor impurity creates free electron (c) P-type
acceptor impurity creates hole, a positive charge carrier.
It is also possible to introduce an impurity lacking an electron as compared with silicon, having
three electrons in the valence shell as compared with four for silicon. In Figure 2.24 (b), this leaves
an empty spot known as a hole, a positive charge carrier. The boron atom tries to bond to four
silicon atoms, but only has three electrons in the valence band. In attempting to form four covalent
bonds the three electrons move around trying to form four bonds. This makes the hole appear to
move. Furthermore, the trivalent atom may borrow an electron from an adjacent (or more distant)
silicon atom to form four covalent bonds. However, this leaves the silicon atom deficient by one
electron. In other words, the hole has moved to an adjacent (or more distant) silicon atom. Holes
reside in the valence band, a level below the conduction band. Doping with an electron acceptor,
an atom which may accept an electron, creates a deficiency of electrons, the same as an excess
of holes. Since holes are positive charge carriers, an electron acceptor dopant is also known as a
P-type dopant. The P-type dopant leaves the semiconductor with an excess of holes, positive charge
carriers. The P-type elements from group IIIA of the periodic table include: boron, aluminum,
gallium, and indium. Boron is used as a P-type dopant for silicon and diamond semiconductors,
while indium is used with germanium.
The marble in a tube analogy to electron conduction in Figure 2.25 relates the movement of
holes with the movement of electrons. The marble represent electrons in a conductor, the tube.
The movement of electrons from left to right as in a wire or N-type semiconductor is explained by
an electron entering the tube at the left forcing the exit of an electron at the right. Conduction
of N-type electrons in inthe conduction band. Compare that with the movement of a hole in the
valence band.
For a hole to enter at the left of Figure 2.25 (b), an electron must be removed. Moving a hole
left to right, the electron must be move right to left. The first electron is ejected from the left end
of the tube so that the hole may move to the right into the tube. The electron is moving in the
opposite direction of the positive hole. As the hole moves farther to the right, electrons must move

50

CHAPTER 2. SOLID-STATE DEVICE THEORY


electron movement

(a)

hole movement

(b)

electron movement

Figure 2.25: Marble in a tube analogy: (a) Electrons move right in the conduction band as electrons
enter tube. (b) Hole moves right in the valence band as electrons move left.
left to accommodate the hole. The hole is the absence of an electron in the valence band due to
P-type doping. It has a localized positive charge. To move the hole in a given direction, the valence
electrons move in the opposite direction.
Electron flow in an N-type semiconductor is similar to electrons moving in a metallic wire. The
N-type dopant atoms will yield electrons available for conduction. These electrons, due to the
dopant are known as majority carriers, for they are in the majority as compared to the very few
holes thermal holes. If an electric field is applied across the N-type semiconductor bar in Figure 2.26
(a), electrons enter the negative (left) end of the bar, traverse the crystal lattice, and exit at right
to the (+) battery terminal.
electron enters

electron exits

crystal lattice

(a) N-type

(b) P-type

Figure 2.26: (a) N-type semiconductor with electrons moving left to right through the crystal lattice.
(b) P-semiconductor with holes moving left to right, which corresponds to electrons moving in the
opposite direction.
Current flow in a P-type semiconductor is a little more difficult to explain. The P-type dopant,
an electron acceptor, yields localized regions of positive charge known as holes. The majority carrier
in a P-type semiconductor is the hole. While holes form at the trivalent dopant atom sites, they may
move about the semiconductor bar. Note that the battery in Figure 2.26 (b) is reversed from (a).
The positive battery terminal is connected to the left end of the P-type bar. Electron flow is out of
the negative battery terminal, through the P-type bar, returning to the positive battery terminal.
An electron leaving the positive (left) end of the semiconductor bar for the positive battery terminal
leaves a hole in the semiconductor, that may move to the right. Holes traverse the crystal lattice
from left to right. At the negative end of the bar an electron from the battery combines with a hole,

2.5.

ELECTRONS AND HOLES

51

neutralizing it. This makes room for another hole to move in at the positive end of the bar toward
the right. Keep in mind that as holes move left to right, that it is actually electrons moving in the
opposite direction that is responsible for the appearant hole movement.
The elements used to produce semiconductors are summarized in Figure 2.27. The oldest elemental semiconductor material is germanium. Though, it is only used to a limited extent today.
Silicon based semiconductors account for about 90% of commercial production of all semiconductors.
Diamond based semiconductors are a research and development activity with considerable potential
at this time. Compound semiconductors not listed include silicon carbide and III-V compounds such
as gallium arsenide. III-VI compound semiconductors include: AlN, GaN, InN, AlP, AlAs, AlSb,
GaP, GaAs, GaSb, InP, InAs, InSb, Alx Ga1x As and Inx Ga1x As. Columns II and VI of periodic
table, not shown in the figure, also form compound semiconductors.
Elemental semiconductors
C(diamond), Si, Ge
B
P-type dopant for C

13

IIIA

5
Boron
10.81

2p1

14
C

IVA

6
Carbon
12.011

2p2

15
N

VA

7
Nitrogen
14.0067

2p3

B, Al, Ga, In
P-type dopant for Si

Al
13 Si
15
14 P
Aluminum
Silicon Phosphorus
26.9815
30.9738
28.0855

Al, Ga, In
P-type dopant for Ge

Ga
31 Ge
33
32 As
Gallium Germanium Arsenic
69.723
72.61
74.92159

3p1

4p1
In

49
Indium
114.82

5p1

3p2

4p2

N, P
N-type dopant for C

3p3

P, As, Sb
N-type dopant for Si, Ge

4p3
Sb
51
Antimony
121.75
5p3

Figure 2.27: Group IIIA P-type dopants, group IV basic semiconductor materials, and group VI
N-type dopants.
The main reason for the inclusion of the IIIA and VIA groups in Figure 2.27 is to show the dopants
used with the group IVA semiconductors. Group IIIA elements are acceptors, P-type dopants, which
accept electrons leaving a hole in the crystal lattice, a positive carrier. Boron is the P-type dopant
for diamond , and most the common dopant for silicon semiconductors. Indium is the P-type dopant
for germanium.
Group VA elements are donors, N-type dopants, yielding a free electron. Nitrogen and Phosphorus are suitable N-type dopants for diamond. Phosphorus and arsenic are the most commonly used
N-type dopants for silicon; though, antimony can be used.
REVIEW:
Intrinsic semiconductor materials, pure to 1 part in 10 billion, are poor conductors.
N-type semiconductor is doped with a pentavalent impurity to create free electrons. Such a
material is conductive. The electron is the majority carrier.
P-type semiconductor, doped with a trivalent impurity, has an abundance of free holes. These

52

CHAPTER 2. SOLID-STATE DEVICE THEORY


are positive charge carriers. The P-type material is conductive. The hole is the majority
carrier
Most semiconductors are based on elements from group IVA of the periodic table, silicon being
the most prevalent. Germanium is all but obsolete. Carbon (diamond) is being developed.
Compound semiconductors such as silicon carbide (group IVA) and gallium arsenide (group
III-V) are widely used.

2.6

The P-N junction

If a block of P-type semiconductor is placed in contact with a block of N-type semiconductor in


Figure 2.28(a), the result is of no value. We have two conductive blocks in contact with each other,
showing no unique properties. The problem is two separate and distinct crystal bodies. The number
of electrons is balanced by the number of protons in both blocks. Thus, neither block has any net
charge.
However, a single semiconductor crystal manufactured with P-type material at one end and Ntype material at the other in Figure 2.28 (b) has some unique properties. The P-type material has
positive majority charge carriers, holes, which are free to move about the crystal lattice. The N-type
material has mobile negative majority carriers, electrons. Near the junction, the N-type material
electrons diffuse across the junction, combining with holes in P-type material. The region of the
P-type material near the junction takes on a net negative charge because of the electrons attracted.
Since electrons departed the N-type region, it takes on a localized positive charge. The thin layer of
the crystal lattice between these charges has been depleted of majority carriers, thus, is known as
the depletion region. It becomes nonconductive intrinsic semiconductor material. In effect, we have
nearly an insulator separating the conductive P an N doped regions.

no charge
separation

electron
(a)

crystal lattice
hole
(b)

intrinsic
charge
separation

Figure 2.28: (a) Blocks of P and N semiconductor in contact have no exploitable properties. (b)
Single crystal doped with P and N type impurities develops a potential barrier.
This separation of charges at the PN junction constitutes a potential barrier. This potential
barrier must be overcome by an external voltage source to make the junction conduct. The formation
of the junction and potential barrier happens during the manufacturing process. The magnitude of

2.6.

THE P-N JUNCTION

53

the potential barrier is a function of the materials used in manufacturing. Silicon PN junctions have
a higher potential barrier than germanium junctions.
In Figure 2.29(a) the battery is arranged so that the negative terminal supplies electrons to
the N-type material. These electrons diffuse toward the junction. The positive terminal removes
electrons from the P-type semiconductor, creating holes that diffuse toward the junction. If the
battery voltage is great enough to overcome the junction potential (0.6V in Si), the N-type electrons
and P-holes combine annihilating each other. This frees up space within the lattice for more carriers
to flow toward the junction. Thus, currents of N-type and P-type majority carriers flow toward
the junction. The recombination at the junction allows a battery current to flow through the PN
junction diode. Such a junction is said to be forward biased.
depletion region
N

electrons
(a)

holes

electrons

holes

(b)

Figure 2.29: (a) Forward battery bias repells carriers toward junction, where recombination results
in battery current. (b) Reverse battery bias attracts carriers toward battery terminals, away from
junction. Depletion region thickness increases. No sustained battery current flows.
If the battery polarity is reversed as in Figure 2.29(b) majority carriers are attracted away from
the junction toward the battery terminals. The positive battery terminal attracts N-type electrons
majority carriers away from the junction. The negative terminal attracts P-type majority carriers,
holes away from the junction. This increases the thickness of the nonconducting depletion region.
There is no recombination of majority carriers; thus, no conduction. This arrangement of battery
polarity is called reverse bias.
The diode schematic symbol is illustrated in Figure 2.30(b) corresponding to the doped semiconductor bar at (a). The diode is a unidirectional device. Electron current only flows in one direction,
against the arrow, corresponding to forward bias. The cathode, bar, of the diode symbol corresponds to N-type semiconductor. The anode, arrow, corresponds to the P-type semiconductor. To
remember this relationship, Not-pointing (bar) on the symbol corresponds to N-type semiconductor.
Pointing (arrow) corresponds to P-type.
If a diode is forward biased as in Figure 2.30(a), current will increase slightly as voltage is
increased from 0 V. In the case of a silicon diode a measurable current flows when the voltage
approaches 0.6 V at (c). As the voltage is increases past 0.6 V, current increases considerably after
the knee. Increasing the voltage well beyond 0.7 V may result in high enough current to destroy the
diode. The forward voltage, VF , is a characteristic of the semiconductor: 0.6 to 0.7 V for silicon, 0.2
V for germanium, a few volts for Light Emitting Diodes (LED). The forward current ranges from a
few mA for point contact diodes to 100 ma for small signal diodes to tens or thousands of amperes
for power diodes.
If the diode is reverse biased, only the leakage of the intrinsic semiconductor flows. This is plotted

54

CHAPTER 2. SOLID-STATE DEVICE THEORY


N

I
mA
electrons

holes

forward
bias V
0.7

reverse bias
(a)
N-type
(not pointing)
(b)

cathode

P-type
(pointing)
anode

breakdown
(c)

Figure 2.30: (a) PN junction, (b) Corresponding diode schematic symbol (c)Silicon Diode I vs V
characteristic curve.
to the left of the origin in Figure 2.30(c). This current will only be as high as 1 A for the most
extreme conditions for silicon small signal diodes. This current does not increase appreciably with
increasing reverse bias until the diode breaks down. At breakdown, the current increases so greatly
that the diode will be destroyed unless a high series resistance limits current. We normally select a
diode with a higher reverse voltage rating than any applied voltage to prevent this. Silicon diodes
are typically available with reverse break down ratings of 50, 100, 200, 400, 800 V and higher. It is
possible to fabricate diodes with a lower rating of a few volts for use as voltage standards.
We previously mentioned that the reverse leakage current of under a A for silicon diodes was due
to conduction of the intrinsic semiconductor. This is the leakage that can be explained by theory.
Thermal energy produces few electron hole pairs, which conduct leakage current until recombination.
In actual practice this predictable current is only part of the leakage current. Much of the leakage
current is due to surface conduction, related to the lack of cleanliness of the semiconductor surface.
Both leakage currents increases with increasing temperature, approaching a A for small silicon
diodes.
For germanium, the leakage current is orders of magnitude higher. Since germanium semiconductors are rarely used today, this is not a problem in practice.
REVIEW:
PN junctions are fabricated from a monocrystalline piece of semiconductor with both a P-type
and N-type region in proximity at a junction.
The transfer of electrons from the N side of the junction to holes annihilated on the P side of
the junction produces a barrier voltage. This is 0.6 to 0.7 V in silicon, and varies with other
semiconductors.
A forward biased PN junction conducts a current once the barrier voltage is overcome. The
external applied potential forces majority carriers toward the junction where recombine takes
place, allowing current flow.

2.7.

JUNCTION DIODES

55

A reverse biased PN junction conducts almost no current. The applied reverse bias attracts
majority carriers away from the junction. This increases the thickness of the nonconducting
depletion region.
Reverse biased PN junctions show a temperature dependent reverse leakage current. This is
less than a A in small silicon diodes.

2.7

Junction diodes

There were some historic crude, but useable semiconductor rectifiers before high purity materials
were available. Ferdinand Braun invented a lead sulfide, PbS, based point contact rectifier in 1874.
Cuprous oxide rectifiers were used as power rectifiers in 1924. The forward voltage drop is 0.2 V.
The linear characteristic curve perhaps is why Cu2 O was used as a rectifier for the AC scale on
DArsonval based multimeters. This diode is also photosensitive.
Selenium oxide rectifiers were used before modern power diode rectifiers became available. This
and the Cu2 O rectifiers were polycrystalline devices. Photoelectric cells were once made from Selenium.
Before the modern semiconductor era, an early diode application was as a radio frequency detector, which recovered audio from a radio signal. The semiconductor was a polycrystalline piece of
the mineral galena, lead sulfide, PbS. A pointed metallic wire known as a cat whisker was brought
in contact with a spot on a crystal within the polycrystalline mineral. (Figure 2.31) The operator
labored to find a sensitive spot on the galena by moving the cat whisker about. Presumably
there were P and N-type spots randomly distributed throughout the crystal due to the variability
of uncontrolled impurities. Less often the mineral iron pyrites, fools gold, was used, as was the
mineral carborundum, silicon carbide, SiC.. Another detector, part of a foxhole radio, consisted of
a sharpened pencil lead bound to a bent safety pin, touching a rusty blue-blade disposable razor
blade. These all required searching for a sensitive spot, easily lost because of vibration.
Replacing the mineral with an N-doped semiconductor (Figure 2.32(a) ) makes the whole surface
sensitive, so that searching for a sensitive spot was no longer required. This device was perfected
by G.W.Pickard in 1906. The pointed metal contact produced a localized P-type region within the
semiconductor. The metal point was fixed in place, and the whole point contact diode encapsulated
in a cylindrical body for mechanical and electrical stability. (Figure 2.32(d) ) Note that the cathode
bar on the schematic corresponds to the bar on the physical package.
Silicon point contact diodes made an important contribution to radar in World War II, detecting
giga-hertz radio frequency echo signals in the radar receiver.. The concept to be made clear is that
the point contact diode preceded the junction diode and modern semiconductors by several decades.
Even to this day, the point contact diode is a practical means of microwave frequency detection
because of its low capacitance. Germanium point contact diodes were once more readily available
than they are today, being preferred for the lower 0.2 V forward voltage in some applications like
self-powered crystal radios. Point contact diodes, though sensitive to a wide bandwidth,, have a low
current capability compared with junction diodes..
Most diodes today are silicon junction diodes. The cross-section in Figure 2.32 looks a bit more
complex than a simple PN junction; though, it is still a PN junction. Starting at the cathode
connection, the N+ indicates this region is heavily doped, having nothing to do with polarity. This
reduces the series resistance of the diode. The N region is lightly doped as indicated by the (-).

56

CHAPTER 2. SOLID-STATE DEVICE THEORY

Figure 2.31: Crystal detector

Anode

Anode

P+

P+

NN+

(a)

(b)
Cathode

Cathode

Anode

Cathode
(c)

(d)

Figure 2.32: Silicon diode cross-section: (a) point contact diode, (b) junction diode, (c) schematic
symbol, (d) small signal diode package.

2.8.

BIPOLAR JUNCTION TRANSISTORS

57

Light doping produces a diode with a higher reverse breakdown voltage, important for high voltage
power rectifier diodes. Lower voltage diodes, even low voltage power rectifiers, would have lower
forward losses with heavier doping. The heaviest level of doping produce zener diodes designed for
a low reverse breakdown voltage. However, heavy doping increases the reverse leakage current. The
P+ region at the anode contact is heavily doped P-type semiconductor, a good contact strategy.
Glass encapsulate small signal junction diodes are capable of 10s to 100s of mA of current. Plastic
or ceramic encapsulated power rectifier diodes handle to 1000s of amperes of current.
REVIEW:
Point contact diodes have superb high frequency characteristics, useable well into the microwave frequencies.
Junction diodes range in size from small signal diodes to power rectifiers capable of 1000s of
amperes.
The level of doping near the junction determines the reverse breakdown voltage. Light doping
produces a high voltage diode. Heavy doping produces a lower breakdown voltage, and increases reverse leakage current. Zener diodes have a lower breakdown voltage because of heavy
doping.

2.8

Bipolar junction transistors

The bipolar transistor (BJT) was named because its operation involves conduction by two carriers:
electrons and holes in the same crystal. The first bipolar transistor was invented at Bell Labs by
William Shockley, Walter Brattain, and John Bardeen so late in 1947 that it was not published until
1948. Thus, many texts differ as to the date of invention. Brattain fabricated a germanium point
contact transistor, bearing some resemblance to a point contact diode. Within a month, Shockley
had a more practical junction transistor, which we describe in following paragraphs. They were
awarded the Nobel Prize in Physics in 1956 for the transistor.
The bipolar junction transistor shown in Figure 2.33(a) is an NPN three layer semiconductor
sandwich with an emitter and collector at the ends, and a base in between. It is as if a third layer
were added to a two layer diode. If this were the only requirement, we would have no more than a
pair of back-to-back diodes. In fact, it is far easier to build a pair of back-to-back diodes. The key
to the fabrication of a bipolar junction transistor is to make the middle layer, the base, as thin as
possible without shorting the outside layers, the emitter and collector. We cannot over emphasize
the importance of the thin base region.
The device in Figure 2.33(a) has a pair of junctions, emitter to base and base to collector, and
two depletion regions.
It is customary to reverse bias the base-collector junction of a bipolar junction transistor as
shown in (Figure 2.33(b). Note that this increases the width of the depletion region. The reverse
bias voltage could be a few volts to tens of volts for most transistors. There is no current flow, save
leakage current, in the collector circuit.
In Figure 2.34(a), a voltage source has been added to the emitter base circuit. Normally we
forward bias the emitter-base junction, overcoming the 0.6 V potential barrier. This is similar to

58

CHAPTER 2. SOLID-STATE DEVICE THEORY


+- P - +
+- - +
+- - +
+- - +
+- - +
+- - +
+- - +
+- - +

emitter

base

+ - P+- +- +- +- +- +- +- -

collector

emitter

+
+
+
+
+
+
+
+

base

collector
-

E
(a)

C
B

E
(b)

C
B

Figure 2.33: (a) NPN junction bipolar transistor. (b) Apply reverse bias to collector base junction.
forward biasing a junction diode. This voltage source needs to exceed 0.6 V for majority carriers
(electrons for NPN) to flow from the emitter into the base becoming minority carriers in the P-type
semiconductor.
If the base region were thick, as in a pair of back-to-back diodes, all the current entering the base
would flow out the base lead. In our NPN transistor example, electrons leaving the emitter for the
base would combine with holes in the base, making room for more holes to be created at the (+)
battery terminal on the base as electrons exit.
However, the base is manufactured thin. A few majority carriers in the emitter, injected as
minority carriers into the base, actually recombine. See Figure 2.34(b). Few electrons injected by
the emitter into the base of an NPN transistor fall into holes. Also, few electrons entering the
base flow directly through the base to the positive battery terminal. Most of the emitter current
of electrons diffuses through the thin base into the collector. Moreover, modulating the small base
current produces a larger change in collector current. If the base voltage falls below approximately
0.6 V for a silicon transistor, the large emitter-collector current ceases to flow.
In Figure 2.35 we take a closer look at the current amplification mechanism. We have an enlarged
view of an NPN junction transistor with emphasis on the thin base region. Though not shown, we
assume that external voltage sources 1) forward bias the emitter-base junction, 2) reverse bias the
base-collector junction. Electrons, majority carriers, enter the emitter from the (-) battery terminal.
The base current flow corresponds to electrons leaving the base terminal for the (+) battery terminal.
This is but a small current compared to the emitter current.
Majority carriers within the N-type emitter are electrons, becoming minority carriers when entering the P-type base. These electrons face four possible fates entering the thin P-type base. A few
at Figure 2.35(a) fall into holes in the base that contributes to base current flow to the (+) battery
terminal. Not shown, holes in the base may diffuse into the emitter and combine with electrons,
contributing to base terminal current. Few at (b) flow on through the base to the (+) battery
terminal as if the base were a resistor. Both (a) and (b) contribute to the very small base current
flow. Base current is typically 1% of emitter or collector current for small signal transistors. The
most of emitter electrons diffuse right through the thin base (c) into the base-collector depletion
region. Note the polarity of the depletion region surrounding the electron at (d). The strong electric
field sweeps the electron rapidly into the collector. The strength of the field is proportional to the
collector battery voltage. Thus 99% of the emitter current flows into the collector. It is controlled

2.8.

BIPOLAR JUNCTION TRANSISTORS

P-

+
+
+
+
+
+
+
+

59

P
-

(a)

C
+

E
E

+
+
+
+
+
+
+
+

C
+

(b)

Figure 2.34: NPN junction bipolar transistor with reverse biased collector-base: (a) Adding forward
bias to base-emitter junction, results in (b) a small base current and large emitter and collector
currents.

(c) (a)
holes
electrons
- + depletion
- + region

(b)
N
emitter

(d)
+
+
+
+
depletion
region
N
P
collector
base

Figure 2.35: Disposition of electrons entering base: (a) Lost due to recombination with base holes.
(b) Flows out base lead. (c) Most diffuse from emitter through thin base into base-collector depletion
region, and (d) are rapidly swept by the strong depletion region electric field into the collector.

60

CHAPTER 2. SOLID-STATE DEVICE THEORY

by the base current, which is 1% of the emitter current. This is a potential current gain of 99, the
ratio of IC /IB , also known as beta, .
This magic, the diffusion of 99% of the emitter carriers through the base, is only possible if the
base is very thin. What would be the fate of the base minority carriers in a base 100 times thicker?
One would expect the recombination rate, electrons falling into holes, to be much higher. Perhaps
99%, instead of 1%, would fall into holes, never getting to the collector. The second point to make
is that the base current may control 99% of the emitter current, only if 99% of the emitter current
diffuses into the collector. If it all flows out the base, no control is possible.
Another feature accounting for passing 99% of the electrons from emitter to collector is that real
bipolar junction transistors use a small heavily doped emitter. The high concentration of emitter
electrons forces many electrons to diffuse into the base. The lower doping concentration in the
base means fewer holes diffuse into the emitter, which would increase the base current. Diffusion of
carriers from emitter to base is strongly favored.
The thin base and the heavily doped emitter help keep the emitter efficiency high, 99% for
example. This correspondes to 100% emitter current splitting between the base as 1% and the
collector as 99%. The emitter efficiency is known as = IC /IE .
Bipolar junction transistors are available as PNP as well as NPN devices. We present a comparison of these two in Figure 2.36. The difference is the polarity of the base emitter diode junctions, as
signified by the direction of the schematic symbol emitter arrow. It points in the same direction as
the anode arrow for a junction diode, against electron current flow. See diode junction, Figure 2.30.
The point of the arrow and bar correspond to P-type and N-type semiconductors, respectively. For
NPN and PNP emitters, the arrow points away and toward the base respectively. There is no
schematic arrow on the collector. However, the base-collector junction is the same polarity as the
base-emitter junction compared to a diode. Note, we speak of diode, not power supply, polarity.

(a)

(b)
E

C
+

E
+

C
-

Figure 2.36: Compare NPN transistor at (a) with the PNP transistor at (b). Note direction of
emitter arrow and supply polarity.
The voltage sources for PNP transistors are reversed compared with an NPN transistors as shown
in Figure 2.36. The base-emitter junction must be forward biased in both cases. The base on a
PNP transistor is biased negative (b) compared with positive (a) for an NPN. In both cases the
base-collector junction is reverse biased. The PNP collector power supply is negative campared with
positive for an NPN transistor.

2.8.

BIPOLAR JUNCTION TRANSISTORS

Collector

61

Collector

Base

NP

(b) Emitter
P+

Base

Emitter
+

N
P base
N collector epitaxial layer

Collector
N+

N+ buried

Base

Emitter (a)

(c)

P substrate

Figure 2.37: Bipolar junction transistor: (a) discrete device cross-section, (b) schematic symbol, (c)
integrated circuit cross-section.
Note that the BJT in Figure 2.37(a) has heavy doping in the emitter as indicated by the N+
notation. The base has a normal P-dopant level. The base is much thinner than the not-to-scale
cross-section shows. The collector is lightly doped as indicated by the N notation. The collector
needs to be lightly doped so that the collector-base junction will have a high breakdown voltage. This
translates into a high allowable collector power supply voltage. Small signal silicon transistors have
a 60-80 V breakdown voltage. Though, it may run to hundreds of volts for high voltage transistors.
The collector also needs to be heavily doped to minimize ohmic losses if the transistor must handle
high current. These contradicting requirements are met by doping the collector more heavily at the
metallic contact area. The collector near the base is lightly doped as compared with the emitter.
The heavy doping in the emitter gives the emitter-base a low approximate 7 V breakdown voltage
in small signal transistors. The heavily doped emitter makes the emitter-base junction have zener
diode like characteristics in reverse bias.
The BJT die, a piece of a sliced and diced semiconductor wafer, is mounted collector down to a
metal case for power transistors. That is, the metal case is electrically connected to the collector.
A small signal die may be encapsulated in epoxy. In power transistors, aluminum bonding wires
connect the base and emitter to package leads. Small signal transistor dies may be mounted directly
to the lead wires. Multiple transistors may be fabricated on a single die called an integrated circuit.
Even the collector may be bonded out to a lead instead of the case. The integrated circuit may
contain internal wiring of the transistors and other integrated components. The integrated BJT
shown in (Figure 2.37) is much thinner than the not to scale drawing. The P+ region isolates
multiple transistors in a single die. An aluminum metalization layer (not shown) interconnects
multiple transistors and other components. The emitter region is heavily doped, N + compared to
the base and collector to improve emitter efficiency.
Discrete PNP transistors are almost as high quality as the NPN counterpart. However, integrated
PNP transistors are not nearly a good as the NPN variety within the same integrated circuit die.
Thus, integrated circuits use the NPN variety as much as possible.
REVIEW:
Bipolar transistors conduct current using both electrons and holes in the same device.

62

CHAPTER 2. SOLID-STATE DEVICE THEORY


Operation of a bipolar transistor as a current amplifier requires that the collector-base junction
be reverse biased and the emitter-base junction be forward biased.
A transistor differs from a pair of back to back diodes in that the base, the center layer, is very
thin. This allows majority carriers from the emitter to diffuse as minority carriers through
the base into the depletion region of the base-collector junction, where the strong electric field
collects them.
Emitter efficiency is improved by heavier doping compared with the collector. Emitter efficiency: = IC /IE , 0.99 for small signal devices
Current gain is =IC /IB , 100 to 300 for small signal transistors.

2.9

Junction field-effect transistors

The field effect transistor was proposed by Julius Lilienfeld in US patents in 1926 and 1933 (1,900,018).
Moreover, Shockley, Brattain, and Bardeen were investigating the field effect transistor in 1947.
Though, the extreme difficulties sidetracked them into inventing the bipolar transistor instead.
Shockleys field effect transistor theory was published in 1952. However, the materials processing
technology was not mature enough until 1960 when John Atalla produced a working device.
A field effect transistor (FET) is a unipolar device, conducting a current using only one kind of
charge carrier. If based on an N-type slab of semiconductor, the carriers are electrons. Conversely,
a P-type based device uses only holes.
At the circuit level, field effect transistor operation is simple. A voltage applied to the gate,
input element, controls the resistance of the channel, the unipolar region between the gate regions.
(Figure 2.38) In an N-channel device, this is a lightly doped N-type slab of silicon with terminals at
the ends. The source and drain terminals are analogous to the emitter and collector, respectively, of
a BJT. In an N-channel device, a heavy P-type region on both sides of the center of the slab serves
as a control electrode, the gate. The gate is analogous to the emitter of a BJT.
Cleanliness is next to godliness applies to the manufacture of field effect transistors. Though
it is possible to make bipolar transistors outside of a clean room, it is a necessity for field effect
transistors. Even in such an environment, manufacture is tricky because of contamination control
issues. The unipolar field effect transistor is conceptually simple, but difficult to manufacture. Most
transistors today are a metal oxide semiconductor variety (later section) of the field effect transistor
contained within integrated circuits. However, discrete JFET devices are available.
A properly biased N-channel junction field effect transistor (JFET) is shown in Figure 2.38. The
gate constitutes a diode junction to the source to drain semiconductor slab. The gate is reverse
biased. If a voltage (or an ohmmeter) were applied between the source and drain, the N-type bar
would conduct in either direction because of the doping. Neither gate nor gate bias is required for
conduction. If a gate junction is formed as shown, conduction can be controlled by the degree of
reverse bias.
Figure 2.39(a) shows the depletion region at the gate junction. This is due to diffusion of
holes from the P=type gate region into the N-type channel, giving the charge separation about the
junction, with a non-conductive depletion region at the junction. The depletion region extends more
deeply into the channel side due to the heavy gate doping and light channel doping.

2.9.

JUNCTION FIELD-EFFECT TRANSISTORS

63

Dra

in
N

Ga

te

Channel

P
So

urc

e
-

Figure 2.38: Junction field effect transistor cross-section.

S N
P-type

S N

D
G

(a)

S N

D
G

(b)

S N

(c)

D
G

(d)

Figure 2.39: N-channel JFET: (a) Depletion at gate diode. (b) Reverse biased gate diode increases
depletion region. (c) Increasing reverse bias enlarges depletion region. (d) Increasing reverse bias
pinches-off the S-D channel.

64

CHAPTER 2. SOLID-STATE DEVICE THEORY

The thickness of the depletion region can be increased Figure 2.39(b) by applying moderate
reverse bias. This increases the resistance of the source to drain channel by narrowing the channel.
Increasing the reverse bias at (c) increases the depletion region, decreases the channel width, and
increases the channel resistance. Increasing the reverse bias VGS at (d) will pinch-off the channel
current. The channel resistance will be very high. This VGS at which pinch-off occurs is VP , the
pinch-off voltage. It is typically a few volts. In summation,the channel resistance can be controlled
by the degree of reverse biasing on the gate.
The source and drain are interchangeable, and the source to drain current may flow in either
direction for low level drain battery voltage ( 0.6 V). That is, the drain battery may be replaced
by a low voltage AC source. For a high drain power supply voltage, to 10s of volts for small signal
devices, the polarity must be as indicated in Figure 2.40(a). This drain power supply, not shown in
previous figures, distorts the depletion region, enlarging it on the drain side of the gate. This is a
more correct representation for common DC drain supply voltages, from a few to tens of volts. As
drain voltage VDS is increased,the gate depletion region expands toward the drain. This increases
the length of the narrow channel, increasing its resistance a little. We say a little because large
resistance changes are due to changing gate bias. Figure 2.40(b) shows the schematic symbol for an
N-channel field effect transistor compared to the silicon cross-section at (a). The gate arrow points
in the same direction as a junction diode. Thepointing arrow and non-pointing bar correspond
to P and N-type semiconductors, respectively.
electron current

to G

P-type
N

D
S

(a)

D
G

(b)

Figure 2.40: N-channel JFET electron current flow from source to drain in (a) cross-section, (b)
schematic symbol.
Figure 2.40 shows a large electron current flow from (-) battery terminal, to FET source, out
the drain, returning to the (+) battery terminal. This current flow may be controlled by varying
the gate voltage. A load in series with the battery sees an amplified version of the changing gate
voltage.
P-channel field effect transistors are also available. The channel is made of P-type material.
The gate is a heavy N-type region. All the voltage sources are reversed in the P-channel circuit
(Figure 2.41) as compared with the more popular N-channel device. Also note, the arrow points out
of the gate of the schematic symbol (b) of the P-channel field effect transistor.
As the positive gate bias voltage is increased, the resistance of the P-channel increases, decreasing
the current flow in the drain circuit.
Discrete devices are manufactured with the cross-section shown in Figure 2.42. The cross-section,
oriented so that it corresponds to the schematic symbol, is upside down with respect to a semiconductor wafer. That is, the gate connections are on the top of the wafer. The gate is heavily doped,
P+ , to diffuse holes well into the channel for a large depletion region. The source and drain connec-

2.9.

JUNCTION FIELD-EFFECT TRANSISTORS

65

to G
S

D
S
G

D
G

N-type

(a)

(b)

Figure 2.41: P-channel JFET: (a) N-type gate, P-type channel, reversed voltage sources compared
with N-channel device. (b) Note reversed gate arrow and voltage sources on schematic.
tions in this N-channel device are heavily doped, N+ to lower connection resistance. However, the
channel surrounding the gate is lightly doped to allow holes from the gate to diffuse deeply into the
channel. That is the N region.

Drain

Drain
Gate

N+
N
P+
Source

N+
Gate

Source

(b)

Source
P+
(a)

Gate

Drain

P+
N
(c) P substrate

Figure 2.42: Junction field effect transistor: (a) Discrete device cross-section, (b) schematic symbol,
(c) integrated circuit device cross-section.
All three FET terminals are available on the top of the die for the integrated circuit version
so that a metalization layer (not shown) can interconnect multiple components. (Figure 2.42(c)
) Integrated circuit FETs are used in analog circuits for the high gate input resistance.. The Nchannel region under the gate must be very thin so that the intrinsic region about the gate can
control and pinch-off the channel. Thus, gate regions on both sides of the channel are not necessary.
The static induction field effect transistor (SIT) is a short channel device with a buried gate.
(Figure 2.43) It is a power device, as opposed to a small signal device. The low gate resistance and
low gate to source capacitance make for a fast switching device. The SIT is capable of hundreds of
amps and thousands of volts. And, is said to be capable of an incredible frequency of 10 gHz.[25]
The Metal semiconductor field effect transistor (MESFET) is similar to a JFET except the gate
is a schottky diode instead of a junction diode. A schottky diode is a metal rectifying contact to a
semiconductor compared with a more common ohmic contact. In Figure 2.44 the the source and
drain are heavily doped (N+ ). The channel is lightly doped (N ). MESFETs are higher speed
than JFETs. The MESET is a depletion mode device, normally on, like a JFET. They are used as

66

CHAPTER 2. SOLID-STATE DEVICE THEORY

Junction field-effect transistor


(static induction type)

Cross-section
Drain

Schematic symbol
N+

Drain

NGate

Gate
+

Source

N+
Source

Figure 2.43: Junction field effect transistor (static induction type): (a) Cross-section, (b) schematic
symbol.

Drain
Gate
(a)

Source

(b)

Source

Gate

N+

N-

Drain
N+

substrate

Figure 2.44: Metal semiconductor field effect transistor (MESFET): (a) schematic symbol, (b) crosssection.

2.10.

INSULATED-GATE FIELD-EFFECT TRANSISTORS (MOSFET)

67

microwave power amplifiers to 30 gHz. MESFETs can be fabricated from silicon, gallium arsenide,
indium phosphide, silicon carbide, and the diamond allotrope of carbon.
REVIEW:
The unipolar junction field effect transistor (FET or JFET) is so called because conduction in
the channel is due to one type of carrier
The JFET source, gate, and drain correspond to the BJTs emitter, base, and collector, respectively.
Application of reverse bias to the gate varies the channel resistance by expanding the gate
diode depletion region.

2.10

Insulated-gate field-effect transistors (MOSFET)

The insulated-gate field-effect transistor (IGFET), also known as the metal oxide field effect transistor (MOSFET), is a derivative of the field effect transistor (FET). Today, most transistors are
of the MOSFET type as components of digital integrated circuits. Though discrete BJTs are more
numerous than discrete MOSFETs. The MOSFET transistor count within an integrated circuit
may approach the hundreds of a million. The dimensions of individual MOSFET devices are under
a micron, decreasing every 18 months. Much larger MOSFETs are capable of switching nearly 100
amperes of current at low voltages; some handle nearly 1000 V at lower currents. These devices
occupy a good fraction of a square centimeter of silicon. MOSFETs find much wider application
than FETs. However, MOSFET power devices are not as widely used as bipolar junction transistors
at this time.
The MOSFET has source, gate, and drain terminals like the FET. However, the gate lead does
not make a direct connection to the silicon compared with the case for the FET. The MOSFET
gate is a metallic or polysilicon layer atop a silicon dioxide insulator. The gate bears a resemblance
to a metal oxide semiconductor (MOS) capacitor in Figure 2.45. When charged, the plates of the
capacitor take on the charge polarity of the respective battery terminals. The lower plate is P-type
silicon from which electrons are repelled by the negative (-) battery terminal toward the oxide, and
attracted by the positive (+) top plate.. This excess of electrons near the oxide creates an inverted
(excess of electrons) channel under the oxide. This channel is also accompanied by a depletion region
isolating the channel from the bulk silicon substrate.
inverted
channel

+
oxide

++++++
oxide

(a)

(b)

- -depletion
- - - -

Figure 2.45: N-channel MOS capacitor: (a) no charge, (b) charged.

68

CHAPTER 2. SOLID-STATE DEVICE THEORY

In Figure 2.46 (a) the MOS capacitor is placed between a pair of N-type diffusions in a P-type
substrate. With no charge on the capacitor, no bias on the gate, the N-type diffusions, the source
and gate, remain electrically isolated.
Source

Gate

Drain

S
+
N
N+

N+

N+

+ + ++++
-- ----

P
(a)

depletion

(b)

inverted channel

D
N+
depletion

Figure 2.46: N-channel MOSFET (enhancement type): (a) 0 V gate bias, (b) positive gate bias.
A positive bias applied to the gate, charges the capacitor (the gate). The gate atop the oxide
takes on a positive charge from the gate bias battery. The P-type substrate below the gate takes
on a negative charge. An inversion region with an excess of electrons forms below the gate oxide.
This region now connects the source and drain N-type regions, forming a continuous N-region from
source to drain. Thus, the MOSFET, like the FET is a unipolar device. One type of charge carrier is
responsible for conduction. This example is an N-channel MOSFET. Conduction of a large current
from source to drain is possible with a voltage applied between these connections. A practical circuit
would have a load in series with the drain battery in Figure 2.46 (b).
The MOSFET described above in Figure 2.46 is known as an enhancement mode MOSFET. The
non-conducting, off, channel is turned on by enhancing the channel below the gate by application of
a bias. This is the most common kind of device. The other kind of MOSFET will not be described
here. See the Insulated-gate field-effect transistor chapter for the depletion mode device.
The MOSFET, like the FET, is a voltage controlled device. A voltage input to the gate controlls
the flow of current from source to drain. The gate does not draw a continuous current. Though, the
gate draws a surge of current to charge the gate capacitance.
The cross-section of an N-channel discrete MOSFET is shown in Figure 2.47 (a). Discrete devices
are usually optimized for high power switching. The N+ indicates that the source and drain are
heavily N-type doped. This minimizes resistive losses in the high current path from source to drain.
The N indicates light doping. The P-region under the gate, between source and drain can be
inverted by application of a positive bias voltage. The doping profile is a across-section, which
may be laid out in a serpentine pattern on the silicon die. This greatly increases the area, and
consequently, the current handling ability.
The MOSFET schematic symbol in Figure 2.47 (b) shows a floating gate, indicating no direct
connection to the silicon substrate. The broken line from source to drain indicates that this device
is off, not conducting, with zero bias on the gate. A normally off MOSFET is an enhancement
mode device. The channel must be enhanced by application of a bias to the gate for conduction.
The pointing end of the substrate arrow corresponds to P-type material, which points toward
an N-type channel, the non-pointing end. This is the symbol for an N-channel MOSFET. The
arrow points in the opposite direction for a P-channel device (not shown). MOSFETs are four
terminal devices: source, gate, drain, and substrate. The substrate is connected to the source in

2.10.

INSULATED-GATE FIELD-EFFECT TRANSISTORS (MOSFET)

69

Drain
N+

Drain

N-

inversion
P

(a)

Gate

Gate
Source

N+
Source

= silicon dioxide
insulator

(b)

Figure 2.47: N-channel MOSFET (enhancement type): (a) Cross-section, (b) schematic symbol.
discrete MOSFETs, making the packaged part a three terminal device. MOSFETs, that are part
of an integrated circuit, have the substrate common to all devices, unless purposely isolated. This
common connection may be bonded out of the die for connection to a ground or power supply bias
voltage.

Drain
N+

Drain
inversion

N-

Gate
P
N+

P
N+

(a)
Gate

Source

Source
= silicon dioxide
insulator

(b)

Figure 2.48: N-channel V-MOS transistor: (a) Cross-section, (b) schematic symbol.
The V-MOS device in (Figure 2.48) is an improved power MOSFET with the doping profile
arranged for lower on-state source to drain resistance. VMOS takes its name from the V-shaped
gate region, which increases the cross-section area of the source-drain path. This minimizes losses
and allows switching of higher levels of power. UMOS, a variation using a U-shaped grove, is more
reproducible in manufacture.
REVIEW:
MOSFETs are unipoar conduction devices, conduction with one type of charge carrier, like a
FET, but unlike a BJT.

70

CHAPTER 2. SOLID-STATE DEVICE THEORY


A MOSFET is a voltage controlled device like a FET. A gate voltage input controls the source
to drain current.
The MOSFET gate draws no continuous current, except leakage. However, a considerable
initial surge of current is required to charge the gate capacitance.

2.11

Thyristors

Thyristors are a broad classification of bipolar-conducting semiconductor devices having four (or
more) alternating N-P-N-P layers. Thyristors include: silicon controlled rectifier (SCR), TRIAC,
gate turn off switch (GTO), silicon controlled switch (SCS), AC diode (DIAC), unijunction transistor
(UJT), programmable unijunction transistor (PUT). Only the SCR is examined in this section;
though the GTO is mentioned.
Shockley proposed the four layer diode thyristor in 1950. It was not realized until years later at
General Electric. SCRs are now available to handle power levels spanning watts to megawatts. The
smallest devices, packaged like small-signal transistors, switch 100s of milliamp at near 100 VAC.
The largest packaged devices are 172 mm in diameter, switching 5600 Amps at 10,000 VAC. The
highest power SCRs may consist of a whole semiconductor wafer several inches in diameter (100s
of mm).

Anode

Anode

P
N

Gate

(a)

Cathode

+
Gate

(b)

Cathode

Figure 2.49: Silicon controlled rectifier (SCR): (a) doping profile, (b) BJT equivalent circuit.
The silicon controlled rectifier is a four layer diode with a gate connection as in Figure 2.49 (a).
When turned on, it conducts like a diode, for one polarity of current. If not triggered on, it is
nonconducting. Operation is explained in terms of the compound connected transistor equivalent in
Figure 2.49 (b). A positive trigger signal is applied between the gate and cathode terminals. This
causes the NPN equivalent transistor to conduct. The collector of the conducting NPN transistor
pulls low, moving the PNP base toward of its collector voltage, which causes the PNP to conduct.
The collector of the conducting PNP pulls high, moving the NPN base in the direction of its collector.
This positive feedback (regeneration) reinforces the NPNs already conducting state. Moreover, the
NPN will now conduct even in the absence of a gate signal. Once an SCR conducts, it continues for
as long as a positive anode voltage is present. For the DC battery shown, this is forever. However,
SCRs are most often used with an alternating current or pulsating DC supply. Conduction ceases

2.11.

THYRISTORS

71

with the expiration of the positive half of the sinewave at the anode. Moreover, most practical SCR
circuits depend on the AC cycle going to zero to cutoff or commutate the SCR.
Figure 2.50 (a) shows the doping profile of an SCR. Note that the cathode, which corresponds
to an equivalent emitter of an NPN transistor is heavily doped as N+ indicates. The anode is
also heavily doped (P+ ). It is the equivalent emitter of a PNP transistor. The two middle layers,
corresponding to base and collector regions of the equivalent transistors, are less heavily doped:
N and P. This profile in high power SCRs may be spread across a whole semiconductor wafer of
substantial diameter.

Anode

schematic symbols

P+
Anode

Anode

NGate
+

(a)
Gate

Gate
Cathode

Cathode

SCR

GTO

(b)

(c)

Cathode

Figure 2.50: Thyristors: (a) Cross-section, (b) silicon controlled rectifier (SCR) symbol, (c) gate
turn-off thyristor (GTO) symbol.
The schematic symbols for an SCR and GTO are shown in Figures 2.50 (b &c). The basic diode
symbol indicates that cathode to anode conduction is unidirectional like a diode. The addition of a
gate lead indicates control of diode conduction. The gate turn off switch (GTO) has bidirectional
arrows about the gate lead, indicating that the conduction can be disabled by a negative pulse, as
well as initiated by a positive pulse.
In addition to the ubiquitous silicon based SCRs, experimental silicon carbide devices have been
produced. Silicon carbide (SiC) operates at higher temperatures, and is more conductive of heat
than any metal, second to diamond. This should allow for either physically smaller or higher power
capable devices.
REVIEW:
SCRs are the most prevalent member of the thyristor four layer diode family.
A positive pulse applied to the gate of an SCR triggers it into conduction. Conduction continues
even if the gate pulse is removed. Conduction only ceases when the anode to cathode voltage
drops to zero.
SCRs are most often used with an AC supply (or pulsating DC) because of the continuous
conduction.
A gate turn off switch (GTO) may be turned off by application of a negative pulse to the gate.

72

CHAPTER 2. SOLID-STATE DEVICE THEORY


SCRs switch megawatts of power, up to 5600 A and 10,000 V.

2.12

Semiconductor manufacturing techniques

The manufacture of only silicon based semiconductors is described in this section; most semiconductors are silicon. Silicon is particularly suitable for integrated circuits because it readily forms an
oxide coating, usefull in patterning integrated components like transistors.
Silicon is the second most common element in the Earths crust in the form of silicon dioxide,
SiO2 , otherwise known as silica sand. Silicon is freed from silicon dioxide by reduction with carbon
in an electric arc furnace
SiO2 + C = CO2 + Si
Such metalurgical grade silicon is suitable for use in silicon steel transformer laminations, but
not nearly pure enough for semiconductor applications. Conversion to the chloride SiCl 4 (or SiHCl3 )
allows purification by fractional distillation. Reduction by ultrapure zinc or magnesium yields sponge
silicon, requiring further purification. Or, thermal decomposition on a hot polycrystalline silicon rod
heater by hydrogen yields ultra pure silicon.
Si + 3HCl
= SiHCl3 + H2
SiHCl3 + H2 = Si + 3HCl2
The polycrystalline silicon is melted in a fused silica crucible heated by an induction heated
graphite suceptor. The graphite heater may alternately be directly driven by a low voltage at high
current. In the Czochralski process, the silicon melt is solidified on to a pencil sized monocrystal
silicon rod of the desired crystal lattice orientation. (Figure 2.51) The rod is rotated and pulled
upward at a rate to encourage the diameter to expand to several inches. Once this diameter is
attained, the boule is automatically pulled at a rate to maintain a constant diameter to a length of a
few feet. Dopants may be added to the crucible melt to create, for example, a P-type semiconductor.
The growing apparatus is enclosed within an inert atmosphere.
The finished boule is ground to a precise final diameter, and the ends trimmed. The boule is
sliced into wafers by an inside diameter diamond saw. The wafers are ground flat and polished. The
wafers could have an N-type epitaxial layer grown atop the wafer by thermal deposition for higher
quality. Wafers at this stage of manufacture are delivered by the silicon wafer manufacturer to the
semiconductor manufacturer.
The processing of semiconductors involves photo lithography, a process for making metal lithographic printing plates by acid etching. The electronics based version of this is the processing of
copper printed circuit boards. This is reviewed in Figure 2.53 as an easy introduction to the photo
lithography involved in semiconductor processing.
We start with a copper foil laminated to an epoxy fiberglass board in Figure 2.53 (a). We also
need positive artwork with black lines corresponding to the copper wiring lines and pads that are
to remain on the finished board. Positive artwork is required because positive acting resist is used.
Though, negative resist is available for both circuit boards and semiconductor processing. At (b)
the liquid positive photo resist is applied to the copper face of the printed circuit board (PCB).
It is allowed to dry and may be baked in an oven. The artwork may be a plastic film positive
reproduction of the original artwork scaled to the required size. The artwork is placed in contact
with the circuit board under a glass plate at (c). The board is exposed to ultraviolet light (d) to
form a latent image of softened photo resist. The artwork is removed (e) and the softened resist
washed away by an alkaline solution (f). The rinsed and dried (baked) circuit board has a hardened

2.12.

SEMICONDUCTOR MANUFACTURING TECHNIQUES

73

lift rod
Si boule
fused silica crucible
graphite suceptor

RF induction coil
Si melt
Figure 2.51: Czochralski monocrystalline silicon growth.

Si boule
cut wafers
diamond blade
driven edge

Figure 2.52: Silicon boule is diamond sawed into wafers.

74

CHAPTER 2. SOLID-STATE DEVICE THEORY

(a) copper PCB

(e) remove artwork

(b) apply photoresist (c) place artwork

(f) develop resist

(g) etch copper

(d) expose

(h) strip resist

Figure 2.53: Processing of copper printed circuit boards is similar to the photo lithographic steps of
semiconductor processing.
resist image atop the copper lines and pads that are to remain after etching. The board is immersed
in the etchant (g) to remove copper not protected by hardened resist. The etched board is rinsed
and the resist removed by a solvent.
The major difference in the patterning of semiconductors is that a silicon dioxide layer atop the
wafer takes the place of the resist during the high temperature processing steps. Though, the resist
is required in low temperature wet processing to pattern the silicon dioxide.
An N-type doped silicon wafer in Figure 2.54 (a) is the starting material in the manufacture
of semiconductor junctions. A silicon dioxide layer (b) is grown atop the wafer in the presence of
oxygen or water vapor at high temperature (over 1000o C in a diffusion furnace. A pool of resist
is applied to the center of the cooled wafer, then spun in a vacuum chuck to evenly distribute the
resist. The baked on resist (c) has a chrome on glass mask applied to the wafer at (d). This mask
contains a pattern of windows which is exposed to ultraviolet light (e).
After the mask is removed in Figure 2.54 (f), the positive resist can be developed (g) in an alkaline
solution, opening windows in the UV softened resist. The purpose of the resist is to protect the
silicon dioxide from the hydrofluoric acid etch (h), leaving only open windows corresponding to the
mask openings. The remaining resist (i) is stripped from the wafer before returning to the diffusion
furnace. The wafer is exposed to a gaseous P-type dopant at high temperature in a diffusion furnace
(j). The dopant only diffuses into the silicon through the openings in the silicon dioxide layer. Each
P-diffusion through an opening produces a PN junction. If diodes were the desired product, the
wafer would be diamond scribed and broken into individual diode chips. However, the whole wafer
may be processed further into bipolar junction transistors.
To convert the diodes into transistors, a small N-type diffusion in the middle of the existing Pregion is required. Repeating the previous steps with a mask having smaller openings accomplishes
this. Though not shown in Figure 2.54 (j), an oxide layer was probably formed in that step during
the P-diffusion. The oxide layer over the P-diffusion is shown in Figure 2.55 (k). Positive photo
resist is applied and dried (l). The chrome on glass emitter mask is applied (m), and UV exposed
(n). The mask is removed (o). The UV softened resist in the emitter opening is removed with an
alkaline solution (p). The exposed silicon dioxide is etched away with hydrofluoric acid (HF) at (q)
After the unexposed resist is stripped from the wafer (r), it is placed in a diffusion furnace

2.12.

SEMICONDUCTOR MANUFACTURING TECHNIQUES

(a) N-type wafer

(b) grow SiO2

(e) expose

(f) remove mask

75

(c) apply photoresist (d) place mask

(g) develop resist

(h) HF etch

BH3

(i) strip resist

(j) P-type diffusion

Figure 2.54: Manufacture of a silicon diode junction.

(k) grow SiO2

(l) apply photoresist

(m) place mask

(n) expose

(o) remove mask

(p) develop resist

(q) HF etch

(r) strip resist

POCl
C

(s) N-type diffusion

(t) metalization

Figure 2.55: Manufacture of a bipolar junction transistor, continuation of Figure 2.54.

76

CHAPTER 2. SOLID-STATE DEVICE THEORY

(Figure 2.55 (s) for high temperature processing. An N-type gaseous dopant, such phosphorus
oxychloride (POCl) diffuses through the small emitter window in the oxide (s). This creates NPN
layers corresponding to the emitter, base, and collector of a BJT. It is important that the N-type
emitter not be driven all the way through the P-type base, shorting the emitter and collector. The
base region between the emitter and collector also needs to be thin so that the transistor has a
usefull . Otherwise, a thick base region could form a pair of diodes rather than a transistor. At
(t) metalization is shown making contact with the transistor regions. This requires a repeat of the
previous steps (not shown here) with a mask for contact openings through the oxide. Another repeat
with another mask defines the metalization pattern atop the oxide and contacting the transistor
regions through the openings.
The metalization could connect numerous transistors and other components into an integrated
circuit. Though, only one transistor is shown. The finished wafer is diamond scribed and broken
into individual dies for packaging. Fine gauge aluminum wire bonds the metalized contacts on the
die to a lead frame, which brings the contacts out of the final package.
REVIEW:
Most semiconductor are based on ultra pure silicon because it forms a glass oxide atop the
wafer. This oxide can be patterned with photo lithography, making complex integrated circuits
possible.
Sausage shaped single crystals of silicon are grown by the Czochralski process, These are
diamond sawed into wafers.
The patterning of silicon wafers by photo lithography is similar to patterning copper printed
circuit boards. Photo resist is applied to the wafer, which is exposed to UV light through a
mask. The resist is developed, then the wafer is etched.
hydrofluoric acid etching opens windows in the protective silicon dioxide atop the wafer.
Exposure to gaseous dopants at high temperature produces semiconductor junctions as defined
by the openings in the silicon dioxide layer.
The photo lithography is repeated for more diffusions, contacts, and metalization.
The metalization may interconnect multiple components into an integrated circuit.

2.13

Superconducting devices

Superconducting devices, though not widely used, have some unique characteristics not available in
standard semiconductor devices. High sensitivity with respect to amplification of electrical signals,
detection of magnetic fields, and detection of light are prized applications. High speed switching is
also possible, though not applied to computers at this time. Conventional superconducting devices
must be cooled to within a few degrees of 0 Kelvin (-273 o C). Though, work is proceeding at
this time on high temperature superconductor based devices, useable at 90 K and below. This is
significant because inexpensive liquid nitrogen may be used for cooling.

2.13.

SUPERCONDUCTING DEVICES

77

Superconductivity: Heike Onnes discovered superconductivity in mercury (Hg) in 1911, for


which he won a Nobel prize. Most metals decrease electrical resistance with decreasing temperature.
Though, most do not decrease to zero resistance as 0 Kelvin is approached. Mercury is unique in
that its resistance abruptly drops to zero at 4.2 K. Superconductors lose all resistance abruptly
when cooled below their critical temperature, Tc A property of superconductivity is no power loss
in conductors. Current may flow in a loop of superconducting wire for thousands of years. Super
conductors include lead (Pb), aluminum, (Al), tin (Sn) and niobium (Nb).
Cooper pair: Lossless conduction in superconductors is not by ordinary electron flow. Electron
flow in normal conductors encounters opposition as collisions with the rigid ionic metal crystal lattice.
Decreasing vibrations of the crystal lattice with decreasing temperature accounts for decreasing
resistance up to a point. Lattice vibrations cease at absolute zero, but not the energy dissipating
collisions of electrons with the lattice. Thus, normal conductors do not lose all resistance at absolute
zero.
Electrons in superconductors form a pair of electrons called a cooper pair, as temperature drops
below the critical temperature at which superconductivity begins. The cooper pair exists because it
is at a lower energy level than unpaired electrons. The electrons are attracted to each other due to
the exchange of phonons, very low energy particles related to vibrations. This cooper pair, quantum
mechanical entity (particle or wave) is not subject to the normal laws of physics. This entity
propagates through the lattice without encountering the metal ions comprising the fixed lattice.
Thus, it dissipates no energy. The quantum mechanical nature of the cooper pair only allows it
to exchange discrete amounts of energy, not continuously variable amounts. An absolute minimum
quantum of energy is acceptable to the cooper pair. If the vibrational energy of the crystal lattice
is less, (due to the low temperature), the cooper pair cannot accept it, cannot be scattered by the
lattice. Thus, under the critical temperature, the cooper pairs flow unimpeded through the lattice.
Josephson junctions: Brian Josephson won a Nobel prize for his 1962 prediction of the Josepheson junction. A Josephson junction is a pair of superconductors bridged by a thin insulator, as in
Figure 2.56 (b), through which electrons can tunnel. The first Josephson junctions were lead superconductors bridged by an insulator. These days a tri-layer of aluminum and niobium is preferred.
Electrons can tunnel through the insulator even with zero voltage applied across the superconductors.
If a voltage is applied across the junction, the current decreases and oscillates at a high frequency
proportional to voltage. The relationship between applied voltage and frequency is so precise that the
standard volt is now defined in terms of Josephson junction oscillation frequency. The Josephson
junction can also serve as a hyper-sensitive detector of low level magnetic fields. It is also very
sensitive to electromagnetic radiation from microwaves to gamma rays.
Josephson transistor: An electrode close to the oxide of the Josephson junction can influence
the junction by capacitive coupling. Such an assembly in Figure 2.56 (b) is Josephson transistor. A
major feature of the Josephson transistor is low power dissipation applicable to high density circuitry,
for example, computers. This transistor is generally part of a more complex superconducting device
like a SQUID or RSFQ.
SQUID: A Superconduction quantum interference device or SQUID is an assembly of Josephson
junctions within a superconducting ring. The DC SQUID is only considered in this discussion. This
device is highly sensitive to low level magnetic fields.
A constant current bias is forced across the ring in parallel with both Josephson junctions in
Figure 2.57. The current divides equally between the two junctions in the absence of an applied
magnetic field and no voltage is developed across across the ring. [4] While any value of Magnetic

78

CHAPTER 2. SOLID-STATE DEVICE THEORY

lead (Pb)

ga

te

lead oxide
(a)

(b)

Figure 2.56: (a) Josephson junction, (b) Josephson transistor.


flux () may be applied to the SQUID, only a quantized value (a multiple of the flux quanta) can
flow through the opening in the superconducting ring.[2] If the applied flux is not an exact multiple of
the flux quanta, the excess flux is cancelled by a circulating current around the ring which produces
a fractional flux quanta. The circulating current will flow in that direction which cancels any excess
flux above a multiple of the flux quanta. It may either add to, or subtract from the applied flux,
up to (1/2) a flux quanta. If the circulating current flows clockwise, the current adds to the top
Josepheson junction and subtracts from the lower one. Changing applied flux linearly causes the
circulating current to vary as a sinusoid.[3] This can be measured as a voltage across the SQUID.
As the applied magnetic field is increased, a voltage pulse may be counted for each increase by a
flux quanta.[19]

+
Iconstant
-

JJ

JJ

counter
V

Figure 2.57: Superconduction quantum interference device (SQUID): Josephson junction pair within
a superconducting ring. A change in flux produces a voltage variation across the JJ pair.
A SQUID is said to be sensitive to 1014 Tesla, It can detect the magnetic field of neural currents
in the brain at 1013 Tesla. Compare this with the 30 x 106 Tesla strength of the Earths magnetic
field.
Rapid single flux quantum (RSFQ): Rather than mimic silicon semiconductor circuits,
RSFQ circuits rely upon new concepts: magnetic flux quantization within a superconductor, and
movement of the flux quanta produces a picosecond quantized voltage pulse. Magnetic flux can only
exist within a section of superconductor quantized in discrete multiples. The lowest flux quanta
allowed is employed. The pulses are switched by Josephson junctions instead of conventional transistors. The superconductors are based on a triple layer of aluminum and niobium with a critical

2.14.

QUANTUM DEVICES

79

temperature of 9.5 K, cooled to 5 K.


RSQFs operate at over 100 gHz with very little power dissipation. Manufacture is simple with
existing photolithographic techniques. Though, operation requires refrigeration down to 5 K . Real
world commercial applications include analog-to-digital and digital to analog converters, toggle flipflops, shift registers, memory, adders, and multipliers.[5]
High temperature superconductors: High temperature superconductors are compounds exhibiting superconductivity above the liquid nitrogen boiling point of 77 K. This is significant because
liquid nitrogen is readily available and inexpensive. Most conventional superconductors are metals;
widely used high temperature superconductors are cuprates, mixed oxides of copper (Cu), for example YBa2 Cu3 O7x , critical temperature, Tc = 90 K . A list of others is available.[23] Most of the
devices described in this section are being developed in high temperature superconductor versions
for less critical applications. Though they do not have the performance of the conventional metal
superconductor devices, the liquid nitrogen cooling is more available.
REVIEW:
Most metals decrease resistance as they approach absolute 0; though, the resistance does
not drop to 0. Superconductors experience a rapid drop to zero resistance at their critical
temperature on being cooled. Typically Tc is within 10 K of absolute zero.
A Cooper pair, electron pair, a quantum mechanical entity, moves unimpeded through the
metal crystal lattice.
Electrons are able to tunnel through a Josephson junction, an insulating gap across a pair of
superconductors.
The addition of a third electrode, or gate, near the junction constitutes a Josephson transistor.
A SQUID, Superconduction quantum interference device, is an highly sensitive detector of
magnetic fields. It counts quantum units of a magnetic field within a superconducting ring.
RSFQ, Rapid single flux quantum is a high speed switching device based on switching the
magnetic quanta existing withing a superconducting loop.
High temperature superconductors, Tc above liquid nitrogen boiling point, may also be used
to build the superconducting devices in this section.

2.14

Quantum devices

Most integrated circuits are digital, based on MOS (CMOS) transistors. Every couple of years since
the late 1960s a geometry shrink has taken place, increasing the circuit density more circuitry at
lower cost in the same space. As of this writing (2006), the MOS transistor gate length is 65-nm
for leading edge production, with 45-nm anticipated within a year. At 65-nm leakage currents were
becoming evident. At 45-nm, heroic innovations were required to minimize this leakage. The end
of shrinkage in MOS transistors is expected at 20- to 30-nm. Though some think that 1- to 2-nm
is the limit. Photolithography, or other lithographic techniques, will continue to improve, providing
ever smaller geometry. However, conventional MOS transistors are not expected to be useable at
these smaller geometries below 20- to 30-nm.

80

CHAPTER 2. SOLID-STATE DEVICE THEORY

Clasical view

Energy

Energy

Energy

Improved photolithography will have to be applied to other than the conventional transistors,
dimensions (under 20- to 30-nm). The objectional MOS leakage currents are due to quantum
mechanical effectselectron tunneling through gate oxide, and the narrow channel. In summary,
quantum mechanical effects are a hindrance to ever smaller conventional MOS transistors. The path
to ever smaller geometry devices involves unique active devices which make practical use of quantum
mechanical principles. As physical geometry becomes very small, electrons may be treated as the
quantum mechanical equivalent: a wave. Devices making use of quantum mechanical principles
include: resonant tunneling diodes, quantum tunneling transistors, metal insulator insulator metal
diodes, and quantum dot transistors.
Quantum tunneling: is the passing of electrons through an insulating barrier which is thin
compared to the de Broglie (page 29) electron wavelength. If the electron wave is large compared
to the barrier, there is a possibility that the wave appears on both sides of the barrier.

Quantum mechanical view

Figure 2.58: Clasical view of an electron surmounting a barrier, or not. Quantum mechanical view
allows an electron to tunnel through a barrier. The probability (green) is related to the barrier
thickness. After Figure 1 [22]
In clasical physics, an electron must have sufficient energy to surmount a barrier. Otherwise, it
recoils from the barrier. (Figure ??. Thinning the barrier increases the probability that the electron
is found on the other side of the barrier. [22]
Tunnel diode: The unqualified term tunnel diode refers to the esaki tunnel diode, an early
quantum device. A reverse biased diode forms a depletion region, an insulating region, between
the conductive anode and cathode. This depletion region is only thin as compared to the electron
wavelength when heavily doped 1000 times the doping of a rectifier diode. With proper biasing,
quantum tunneling is possible. See (page 138)for details.
RTD, resonant tunneling diode: This is a quantum device not to be confused with the Esaki
tunnel diode, (page 138) , a conventional heavily doped bipolar semiconductor. Electrons tunnel
through two barriers separated by a well in flowing source to drain in a resonant tunneling diode.
Tunneling is also known as quantum mechanical tunneling. The flow of electrons is controlled by
diode bias. This matches the energy levels of the electrons in the source to the quantized level in
the well so that electrons can tunnel through the barriers. The energy level in the well is quantized
because the well is small. When the energy levels are equal, a resonance occurs, allowing electron
flow through the barriers as shown in Figure 2.59 (b). No bias or too much bias, in Figures 2.59 (a)
and (c) respectively, yields an energy mismatch between the source and the well, and no conduction.
As bias is increased from zero across the RTD, the current increases and then decreases, corresponding to off, on, and off states. This makes simplification of conventional transistor circuits
possible by substituting a pair of RTDs for two transistors. For example, two back-to-back RTDs

InGaAs

InGaAs

GaAs

(a)

10

20

30

barrier
energy
level

Energy eV

barrier
well

AlAs

InGaAs

AlAs

GaAs

81
Energy eV

QUANTUM DEVICES
Energy eV

2.14.

0
source

40 nm

10

(b)

drain
20

30

40 nm

10

20

30

40 nm

(c)

Figure 2.59: Resonant tunneling diode (RTD): (a) No bias, source and well energy levels not matched,
no conduction. (b) Small bias causes matched energy levels (resonance); conduction results. (c)
Further bias mismatches energy levels, decreasing conduction.
and a transistor form a memory cell, using fewer components, less area and power compared with a
conventional circuit. The potential application of RTDs is to reduce the component count, area, and
power dissipation of conventional transistor circuits by replacing some, though not all, transistors.
[11] RTDs have been shown to oscillate up to 712 gHz. [8]
Double-layer tunneling transistor: The Deltt, otherwise known as the Double-layer tunneling
transistor is constructed of a pair of conductive wells separated by an insulator or high band gap
semiconductor. (Figure 2.60) The wells are so thin that electrons are confined to two dimensions.
These are known as quantum wells. A pair of these quantum wells are insulated by a thin GaAlAs,
high band gap (does not easily conduct) layer. Electrons can tunnel through the insulating layer
if the electrons in the two quantum wells have the same momentum and energy. The wells are so
thin that the electron may be treated as a wave the quantum mechanical duality of particles and
waves. The top and optional bottom control gates may be adjusted to equalize the energy levels
(resonance) of the electrons to allow conduction from source to drain. Figure 2.60, barrier diagram
red bars show unequal energy levels in the wells, an off-state condition. Proper biasing of the
gates equalizes the energy levels of electrons in the wells, the on-state condition. The bars would
be at the same level in the energy level diagram.
If gate bias is increased beyond that required for tunneling, the energy levels in the quantum
wells no longer match, tunneling is inhibited, source to drain current decreases. To summarize,
increasing gate bias from zero results in on, off, on conditions. This allows a pair of Deltts to be
stacked in the manner of a CMOS complementary pair; though, different p- and n-type transistors
are not required. Power supply voltage is about 100 mV. Experimental Deltts have been produced
which operate near 4.2 K, 77 K, and 0o C. Room temperature versions are expected.[11] [13] [20]
MIIM diode: The metal-insulator-insulator-metal (MIIM) diode is a quantum tunneling device,
not based on semiconductors. See MIIM diode section Figure 2.61. The insulator layers must be
thin compared to the de Broglie (page 29) electron wavelength, for quantum tunneling to be possible.
For diode action, there must be a prefered tunneling direction, resulting in a sharp bend in the diode
forward characteristic curve. The MIIM diode has a sharper forward curve than the metal insulator
metal (MIM) diode, not considered here.
The energy levels of M1 and M2 are equal in no bias Figure ?? is increased, a quantum well,
an area where electrons may exist, is formed between the insulators. Electrons may pass through
insulator I1 if M1 is bised at the same energy level as the quantum well. A simple explanation is
that the distance through the insulators is shorter. A longer explanation is that as bias increases,
the probability of the electron wave everlapping from M1 to the quantum well increases. For a more

82

CHAPTER 2. SOLID-STATE DEVICE THEORY

GaAs

top quantum well

well

tunneling

barrier
well
depletion
barrier
diagram

top quantum well


contact (drain)

top depletion
gate

bottom quantum well


contact (drain)
top gate

depletion

GaAlAs

insulator

GaAs
bottom quantum well
bottom gate (optional)

bottom
depletion gate

quantum
well

Energy

I1

Energy

M1

Energy

Figure 2.60: Double-layer tunneling transistor (Deltt) is composed of two electron containing wells
separated by a nonconducting barrier. The gate voltages may be adjusted so that the energy and
momentum of the electrons in the wells are equal which permits electrons to tunnel through the
nonconductive barrier. (The energy levels are shown as unequal in the barrier diagram.)

I2
M2

M1 I1 I2 M2
Distance

MIIM diode
section

No bias

M1 I2 I1 M2
Distance
Forward bias

M1 I1 I2

M2

Distance
Reverse bias

Figure 2.61: Metal insulator insulator metal (MIIM) diode: Cross section of diode. Energy levels
for no bias, forward bias, and reverse bias. After Figure 1 [21].

2.14.

QUANTUM DEVICES

83

detailed explanation see Phiar Corp. [21]


MIIM devices operate at higher frequencies (3.7 THz) than microwave transistors. [16] The
addition of a third electrode to a MIIM diode produces a transistor.
Quantum dot transistor: An isolated conductor may take on a charge, measured in coulombs
for large objects. For a nano-scale isolated conductor known as a quantum dot, the charge is measured
in electrons. A quantum dot of 1- to 3-nm may take on an incremental charge of a single electron.
This is the basis of the quantum dot transistor, also known as a single electron transistor.
A quantum dot placed atop a thin insulator over an electron rich source is known as a single
electron box. (Figure 2.62 (a)) The energy required to transfer an electron is related to the size of
the dot and the number of electrons already on the dot.
A gate electrode above the quantum dot can adjust the energy level of the dot so that quantum
mechanical tunneling of an electron (as a wave) from the source through the insulator is possible.
(Figure 2.62 (b)) Thus, a single electron may tunnel to the dot.

+
gate

+++

quantum
dot

tunneling

tunnel
barrier

--+++

source

drain

---

tunnel
barrier

source
(a)

(b)

(c)

Figure 2.62: (a) Single electron box, an isolated quantum dot separated from an electron source
by an insulator. (b) Positive charge on the gate polarizes quantum dot, tunneling an electron from
the source to the dot. (c) Quantum transistor: channel is replaced by quantum dot surrounded by
tunneling barrier.
If the quantum dot is surrounded by a tunnel barrier and embedded between the source and
drain of a conventional FET, as in Figure 2.62 (c) , the charge on the dot can modulate the flow
of electrons from source to drain. As gate voltage increases, the source to drain current increases,
up to a point. A further increase in gate voltage decreases drain current. This is similar to the
behavior of the RTD and Deltt resonant devices. Only one kind of transistor is required to build a
complementary logic gate.[11]
Single electron transistor: If a pair of conductors, superconductors, or semiconductors are
separated by a pair tunnel barriers (insulator), surrounding a tiny conductive island, like a quantum
dot, the flow of a single charge (a Cooper pair for superconductors) may be controlled by a gate.
This is single electron transistor similar to Figure 2.62 (c). Increasing the positive charge on the
gate, allows an electron to tunnel to the island. If it is sufficiently small, the low capacitance will
cause the dot potential to rise substantially due to the single electron. No more electrons can tunnel
to the island due the electron charge. This is known at the coulomb blockade. The electron which
tunneled to the island, can tunnel to the drain.
Single electron transistors operate at near absolute zero. The exception is the graphene single
electron transistor, having a graphene island. They are all experimental devices.

84

CHAPTER 2. SOLID-STATE DEVICE THEORY

Graphene transistor: Graphite, an allotrope of carbon, does not have the rigid interlocking
crystalline structure of diamond. None the less, it has a crystalline structure one atom thick, a
so called two-dimensional structure. A graphite is a three-dimensional crystal. However, it cleaves
into thin sheets. Experimenters, taking this to the extreme, produce micron sized specks as thin
as a single atom known as graphene. (Figure 2.63 (a)) These membranes have unique electronic
properties. Highly conductive, conduction is by either electrons or holes, without doping of any
kind. [12]
Graphene sheets may be cut into transistor structures by lithographic techniques. The transistors
bear some resemblance to a MOSFET. A gate capacitively coupled to a graphene channel controls
conduction.
As silicon transistors scale to smaller sizes, leakage increases along with power dissipation. And
they get smaller every couple of years. Graphene transistors dissipate little power. And, they switch
at high speed. Graphene might be a replacement for silicon someday.
Graphene can be fashioned into devices as small as sixty atoms wide. Graphene quantum dots
within a transistor this small serve as single electron transistors. Previous single electron transistors
fashioned from either superconductors or conventional semiconductors operate near absolute zero.
Graphene single electron transistors uniquely function at room temperature.[24]
Graphene transistors are laboratory curiosities at this time. If they are to go into production
two decades from now, graphene wafers must be produced. The first step, production of graphene
by chemical vapor deposition (CVD) has been accomplished on an experimental scale. Though, no
wafers are available to date.

(a)

(b)

Figure 2.63: (a) Graphene: A single sheet of the graphite allotrope of carbon. The atoms are
arranged in a hexagonal pattern with a carbon at each intersection. (b) Carbon nanotube: A
rolled-up sheet of graphene.
Carbon nanotube transistor: If a 2-D sheet of graphene is rolled, the resulting 1-D structure
is known as a carbon nanotube. (Figure 2.63 (b)) The reason to treat it as 1-dimensional is that it is
highly conductive. Electrons traverse the carbon nanotube with being scattered by a crystal lattice.
Resistance in normal metals is caused by scattering of electrons by the metallic crystalline lattice. If
electrons avoid this scattering, conduction is said to be by ballistic transport. Both metallic (acting)
and semiconducting carbon nanotubes have been produced. [6]
Field effect transistors may be fashioned from a carbon nanotubes by depositing source and drain
contacts on the ends, and capacitively coupling a gate to the nanotube between the contacts. Both
p- and n-type transistors have been fabricated. Why the interest in carbon nanotube transistors?
Nanotube semiconductors are Smaller, faster, lower power compared with silicon transistors. [7]

2.14.

QUANTUM DEVICES

85

Spintronics: Conventional semiconductors control the flow of electron charge, current. Digital
states are represented by on or off flow of current. As semiconductors become more dense
with the move to smaller geometry, the power that must be dissipated as heat increases to the
point that it is difficult to remove. Electrons have properties other than charge such as spin. A
tentative explanation of electron spin is the rotation of distributed electron charge about the spin
axis, analogous to diurnal rotation of the Earth. The loops of current created by charge movement,
form a magnetic field. However, the electron is more like a point charge than a distributed charge,
Thus, the rotating distributed charge analogy is not a correct explanation of spin. Electron spin
may have one of two states: up or down which may represent digital states. More precisely the spin
(ms ) quantum number may be 1/2 the angular momentum (l) quantum number. [1]
Controlling electron spin instead of charge flow considerably reduces power dissipation and increases switching speed. Spintronics, an acronym for SPIN TRansport electrONICS, is not widely
applied because of the difficulty of generating, controlling, and sensing electron spin. However, high
density, non-volatile magnetic spin memory is in production using modified semiconductor processes.
This is related to the spin valve magnetic read head used in computer harddisk drives, not mentioned
further here.
A simple magnetic tunnel junction (MTJ) is shown in Figure 2.64 (a), consisting of a pair
of ferromagnetic, strong magnetic properties like iron (Fe), layers separated by a thin insulator.
Electrons can tunnel through a sufficiently thin insulator due to the quantum mechanical properties
of electrons the wave nature of electrons. The current flow through the MTJ is a function of the
magnetization, spin polarity, of the ferromagnetic layers. The resistance of the MTJ is low if the
magnetic spin of the top layer is in the same direction (polarity) as the bottom layer. If the magnetic
spins of the two layers oppose, the resistance is higher. [9]

contact
ferromagnet
tunneling
insulator
ferromagnet
contact
antiferromagnet
contact
(a)

(b)

Figure 2.64: (a) Magnetic tunnel junction (MTJ): Pair of ferromagnetic layers separated by a thin
insulator. The resistance varies with the magnetization polarity of the top layer (b) Antiferromagnetic bias magnet and pinned bottom ferromagnetic layer increases resistance sensitivity to changes
in polarity of the top ferromagnetic layer. Adapted from [9] Figure 3.
The change in resistance can be enhanced by the addition of an antiferromagnet, material having
spins aligned but opposing, below the bottom layer in Figure 2.64 (b). This bias magnet pins the
lower ferromagnetic layer spin to a single unchanging polarity. The top layer magnetization (spin)

86

CHAPTER 2. SOLID-STATE DEVICE THEORY

may be flipped to represent data by the application of an external magnetic field not shown in the
figure. The pinned layer is not affected by external magnetic fields. Again, the MTJ resistance
is lowest when the spin of the top ferromagnetic layer is the same sense as the bottom pinned
ferromagnetic layer. [9]
The MTJ may be improved further by splitting the pinned ferromagnetic layer into two layers
separated by a buffer layer in Figure 2.65 (a). This isolates the top layer. The bottom ferromagnetic
layer is pinned by the antiferromagnet as in the previous figure. The ferromagnetic layer atop the
buffer is attracted by the bottom ferromagnetic layer. Opposites attract. Thus, the spin polarity
of the additional layer is opposite of that in the bottom layer due to attraction. The bottom and
middle ferromagnetic layers remain fixed. The top ferromagnetic layer may be set to either spin
polarity by high currents in proximate conductors (not shown). This is how data are stored. Data
are read out by the difference in current flow through the tunnel junction. Resistance is lowest if
the layers on both sides of the insulting layer are of the same spin. [9]

data

top contact
ferromagnet

pinned layers

tunneling
insulator
ferromagnet
coupling layer
ferromagnet
antiferromagnet

(a)

bottom contact

(b)

Figure 2.65: (a)Splitting the pinned ferromagnetic layer of Figure 2.64 (b) by a buffer layer improves
stability and isolates the top ferromagnetic unpinned layer. Data are stored in the top ferromagnetic
layer based on spin polarity (b) MTJ cell embedded in read lines of a semiconductor die one of
many MTJs. Adapted from [10]
An array of magnetic tunnel junctions may be embedded in a silicon wafer with conductors
connecting the top and bottom terminals for reading data bits from the MTJs with conventional
CMOS circuitry. One such MTJ is shown in Figure 2.65 (b) with the read conductors. Not shown,
another crossed array of conductors carrying heavy write currents switch the magnetic spin of the
top ferromagnetic layer to store data. A current is applied to one of many X conductors and a
Y conductor. One MTJ in the array is magnetized under the conductors cross-over. Data are
read out by sensing the MTJ current with conventional silicon semiconductor circuitry. [10]
The main reason for interest in magnetic tunnel junction memory is that it is nonvolatile. It does
not lose data when powered off. Other types of nonvolatile memory are capable of only limited
storage cycles. MTJ memory is also higher speed than most semiconductor memory types. It is now
(2006) a commercial product. [18]

2.15.

SEMICONDUCTOR DEVICES IN SPICE

87

Not a commercial product, or even a laboratory device, is the theoretical spin transistor which
might one day make spin logic gates possible. The spin transistor is a derivative of the theoretical
spin diode.
It has been known for some time that electrons flowing through a cobalt-iron ferromagnet become
spin polarized. The ferromagnet acts as a filter passing electrons of one spin preferentially. These
electrons may flow into an adjacent nonmagnetic conductor (or semiconductor) retaining the spin
polarization for a short time, nano-seconds. Though, spin polarized electrons may propagate a
considerable distance compared with semiconductor dimensions. The spin polarized electrons may
be detected by a nickel-iron ferromagnetic layer adjacent to the semiconductor. [1] [15]
It has also been shown that electron spin polarization occurs when circularly polarized light
illuminates some semiconductor materials. Thus, it should be possible to inject spin polarized
electrons into a semiconductor diode or transistor. The interest in spin based transistors and gates
is because of the non-dissipative nature of spin propagation, compared with dissipative charge flow.
As conventional semiconductors are scaled down in size, power dissipation increases. At some point
the scaling down will no longer be practical. Researchers are looking for a replacement for the
conventional charge flow based transistor. That device may be based on spintronics. [14]
REVIEW:
As MOS gate oxide thins with each generation of smaller transistors, excessive gate leakage
causes unacceptable power dissipation and heating. The limit of scaling down conventional
semiconductor geometry is within sight.
Resonant tunneling diode (RTD): Quantum mechanical effects, which degrade conventional
semiconductors, are employed in the RTD. The flow of electrons through a sufficiently thin
insulator, is by the wave nature of the electron particle wave duality. The RTD functions as
an amplifier.
Double layer tunneling transistor (Deltt): The Deltt is a transistor version of the RTD. Gate
bias controls the ability of electrons to tunnel through a thin insulator from one quantum well
to another (source to drain).
Quantum dot transistor: A quantum dot, capable of holding a charge, is surrounded by a
thin tunnel barrier replacing the gate of a conventional FET. The charge on the quantum dot
controls source to drain current flow.
Spintronics: Electrons have two basic properties: charge and spin. Conventional electronic
devices control the flow of charge, dissipating energy. Spintronic devices manipulate electron
spin, a propagative, non-dissipative process.

2.15

Semiconductor devices in SPICE

The SPICE (simulation program, integrated circuit emphesis) electronic simulation program provides circuit elements and models for semiconductors. The SPICE element names begin with d, q,
j, or m correspond to diode, BJT, JFET and MOSFET elements, respectively. These elements are
accompanied by corresponding models These models have extensive lists of parameters describing
the device. Though, we do not list them here. In this section we provide a very brief listing of

88

CHAPTER 2. SOLID-STATE DEVICE THEORY

simple spice models for semiconductors, just enough to get started. For more details on models and
an extensive list of model parameters see Kuphaldt. [17] This reference also gives instructions on
using SPICE.
Diode: The diode statement begins with a diode element name which must begin with d
plus optional characters. Some example diode element names include: d1, d2, dtest, da, db, d101,
etc. Two node numbers specify the connection of the anode and cathode, respectively, to other
components. The node numbers are followed by a model name, referring to a .model statement.
The model statement line begins with .model, followed by the model name matching one or
more diode statements. Next is a d indicating that a diode is being modeled. The remainder of the
model statement is a list of optional diode parameters of the form ParameterName=ParameterValue.
None are shown in the example below. For a list, see reference, diodes. [17]
General form: d[name] [anode] [cathode] [model]
.model [modelname] d ( [parmtr1=x] [parmtr2=y] . . .)
Example:

d1 1 2 mod1
.model mod1 d

Models for specific diode part numbers are often furnished by the semiconductor diode manufacturer. These models include parameters. Otherwise, the parameters default to so called default
values, as in the example.
BJT, bipolar junction transistor: The BJT element statement begins with an element name
which must begin with q with associated circuit symbol designator characters, example: q1, q2,
qa, qgood. The BJT node numbers (connections) identify the wiring of the collector, base, emitter
respectively. A model name following the node numbers is associated with a model statement.
General form: q[name] [collector] [base] [emitter] [model]
.model [modelname] [npn or pnp] ([parmtr1=x] . . .)
Example:
Example:

q1 2 3 0 mod1
.model mod1 pnp
q2 7 8 9 q2n090
.model q2n090 npn ( bf=75 )

The model statement begins with .model, followed by the model name, followed by one of
npn or pnp. The optional list of parameters follows, and may continue for a few lines beginning
with line continuation symbol +, plus. Shown above is the forward parameter set to 75 for
the hypothetical q2n090 model. Detailed transistor models are often available from semiconductor
manufacturers.
FET, field effect transistor The field effect transistor element statement begins with an
element name beginning with j for JFET associated with some unique characters, example: j101,
j2b, jalpha, etc. The node numbers follow for the drain, gate and source terminals, respectively.
The node numbers define connectivity to other circuit components. Finally, a model name indicates
the JFET model to use.
General form: j[name] [drain] [gate] [source] [model]
.model [modelname] [njf or pjf] ( [parmtr1=x] . . .)
Example:

j1 2 3 0 mod1

2.15.

SEMICONDUCTOR DEVICES IN SPICE

89

.model mod1 pjf


j3 4 5 0 mod2
.model mod2 njf ( vto=-4.0 )
The .model in the JFET model statement is followed by the model name to identify this
model to the JFET element statement(s) using it. Following the model name is either pjf or njf for
p-channel or n-channel JFETs respectively. A long list of JFET parameters may follow. We only
show how to set Vp , pinch off voltage, to -4.0 V for an n-channel JFET model. Otherwise, this vto
parameter defaults to -2.5 V or 2.5V for n-channel or p-channel devices, respectively.
MOSFET, metal oxide field effect transistor The MOSFET element name must begin
with m, and is the first word in the element statement. Following are the four node numbers
for the drain, gate, source, and substrate, respectively. Next is the model name. Note that the
source and substrate are both connected to the same node 0 in the example. Discrete MOSFETs
are packaged as three terminal devices, the source and substrate are the same physical terminal.
Integrated MOSFETs are four terminal devices; the substrate is a fourth terminal. Integrated
MOSFETs may have numerous devices sharing the same substrate, separate from the sources.
Though, the sources might still be connected to the common substrate.
General form: m[name] [drain] [gate] [source] [substrate] [model]
.model [modelname] [nmos or pmos] ( [parmtr1=x] . . . )
Example:

m1 2 3
m5 5 6
.model
.model

0 0 mod1
0 0 mod4
mod1 pmos
mod4 nmos ( vto=1 )

The MOSFET model statement begins with .model followed by the model name followed
by either pmos or nmos. Optional MOSFET model parameters follow. The list of possible
parameters is long. See Volume 5, MOSFET for details. [17] MOSFET manufacturers provide
detailed models. Otherwise, defaults are in effect.
The bare minimum semiconductor SPICE information is provided in this section. The models
shown here allow simulation of basic circuits. In particular, these models do not account for high
speed or high frequency operation. Simulations are shown in the Volume 5 Chapter 7, Using SPICE
....
REVIEW:
Semiconductors may be computer simulated with SPICE.
SPICE provides element statements and models for the diode, BJT, JFET, and MOSFET.

Contributors
Contributors to this chapter are listed in chronological order of their contributions, from most recent
to first. See Appendix 2 (Contributor List) for dates and contact information.
Maciej Noszczyski (December 2003): Corrected spelling of Niels Bohrs name.
Bill Heath (September 2002): Pointed out error in illustration of carbon atom the nucleus
was shown with seven protons instead of six.

90

BIBLIOGRAPHY

Bibliography
[1] David D. Awschalom, Michael E. Flatt, Nitin Samarth, Spintronics, Scientific American,
June 2002 at http://www.sciam.com
[2] John Bland, The Fluxoid in A Mssbauer Spectroscopy and Magnetometry Study of Magnetic Multilayers and Oxides, Oliver Lodge Laboratory, Department of Physics, University of
Liverpool, 2002, at http://www.cmp.liv.ac.uk/frink/thesis/thesis/node45.html
[3] John Bland, Superconducting Quantum Interference Device (SQUID) in A Mssbauer Spectroscopy and Magnetometry Study of Magnetic Multilayers and Oxides,
Oliver Lodge Laboratory, Department of Physics, University of Liverpool, 2002, at
http://www.cmp.liv.ac.uk/frink/thesis/thesis/node48.html
[4] John Bland, SQUID Magnetometer in A Mssbauer Spectroscopy and Magnetometry Study
of Magnetic Multilayers and Oxides, Oliver Lodge Laboratory, Department of Physics, University of Liverpool, 2002, at http://www.cmp.liv.ac.uk/frink/thesis/thesis/node48.html
[5] Darren K. Brock, RSFQ Technology: Circuits and Systems, Hypres, Inc., NY, at
http://www.hypres.com/papers/Brock-RSFQ-CirSys-Rev01.pdf
[6] Matthew Broersma , Nanotubes break semiconducting record, Cnet News, December 19,
2003, at http://news.com.com/2100-1006-5129761.html
[7] Carbon Nanotube Transistor, Physics News
http://www.aip.org/mgr/png/html/tubefet.htm

Graphics,

May

13,

1998,

at

[8] E. R. Brown, C. D. Parker, Resonant Tunnel Diodes as Submillimetre-Wave


Sources,
Philosophical Transactions:
Mathematical,
Physical and Engineering Sciences, Vol. 354, No. 1717, The Current Status of Semiconductor Tunnelling Devices (Oct. 15, 1996), pp. 2365-2381 at http://links.jstor.org/
sici?sici=1364-503X(19961015)354%3A1717%3C2365%3ARTDASS%3E2.0.CO%3B2-Q
[9] W. J. Gallagher, S. S. P. Parkin, Development of the magnetic tunnel junction MRAM at IBM: From first junctions to a 16-Mb MRAM demonstrator chip,
IBM Research and Development, Spintronics, Volume 50, Number 1, 2006, at
http://www.research.ibm.com/journal/rd/501/gallagher.html
[10] IBM, Infineon Develop Most Advanced MRAM Technology to Date, IBM Research, at
http://domino.research.ibm.com/comm/pr.nsf/pages/news.20030610 mram.html
[11] Linda Geppert Quantum Transistors: toward nanoectronic, IEEE Spectrum, September 2000,
at http://www.ece.osu.edu/~berger/press/quan0900.pdf
[12] A. K. Geim1 and K. S. Novoselov1 , The rise of graphene, Nature Materials, 6, 2007, at
http://www.nature.com/nmat/journal/v6/n3/full/nmat1849.html
[13] Ilan Greenberg, Transistor Technology Takes a Quantum Leap, Wired News, December 8,
1997, at http://www.wired.com/news/technology/0,1282,8994,00.html

BIBLIOGRAPHY

91

[14] R. Colin Johnson, Spintronics approach advances toward live chips, EE Times, 11/06/2006,
at http://www.eetimes.com/showArticle.jhtml?articleID=193500309
[15] R. Colin Johnson U. of Delaware researchers edge closer to spintronics, EE Times,
07/26/2007, at http://www.eetimes.com/news/design/showArticle.jhtml?articleID=201201400
[16] R. Colin Johnson, Can metal-insulator electronics do it better, sans semiconductors?
http://www.eetimes.com/showArticle.jhtml?articleID=201200024
[17] Tony R. Kuphaldt, Lessons in Electricity, Reference, Vol. 5,
http://www.ibiblio.org/obp/electricCircuits/Ref/spice.html

Ch 7,

2007 at

[18] Tom
Lee,
Is
nonvolatile
MRAM
right
for
your
consumer
bedded
device
application?
,
Freescale
Semiconductor
http://www.acumeninfo.com/subscriber/article/getArticle.jhtml?
articleId=197006965
[19] HyperPhysics,
SQUID
Magnetometer,
HyperPhysics
http://hyperphysics.phy-astr.gsu.edu/hbase/solids/squid.html

emat

at

[20] Phillip F. Schewe, Ben Stein, A Quantum Tunneling Transistor, Physics Nessw Update,
Number 357, February 4, 1998, at http://www.aip.org/pnu/1998/physnews.357.htm
[21] Why MIIM?, Phiar Corporation, at http://www.phiar.com/whyMIIM.php4
[22] What
is
Quantum
Tunneling?,
http://www.phiar.com/whatQuantum.php4

Phiar

Corporation,

at

[23] Oxford University, Theory, Superconductor Synthesis, Oxford University, 1996, at


http://www.chem.ox.ac.uk/vrchemistry/super/theory.htm
[24] John Walko, Graphene transistor to rival silicon, say researchers, EE Times Europe,
03/02/2007,
at
http://www.eetimes.com/news/design/showArticle.jhtml?
articleID=197700700
[25] Ying-Yu Tzou,Power Electronics: An Introduction, Institute of Control Engineering, National
Chiao Tung University, at http://pemclab.cn.nctu.edu.tw/peclub/w3cnotes

92

BIBLIOGRAPHY

Chapter 3

DIODES AND RECTIFIERS


Contents
3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8
3.9

Meter check of a diode . . . . . . . . . . . . .


Diode ratings . . . . . . . . . . . . . . . . . .
Rectifier circuits . . . . . . . . . . . . . . . .
Peak detector . . . . . . . . . . . . . . . . . .
Clipper circuits . . . . . . . . . . . . . . . . .
Clamper circuits . . . . . . . . . . . . . . . .
Voltage multipliers . . . . . . . . . . . . . . .
Inductor commutating circuits . . . . . . . .
Diode switching circuits . . . . . . . . . . . .
3.9.1
Logic . . . . . . . . . . . . . . . . . . . . . .
3.9.2
Analog switch . . . . . . . . . . . . . . . . .
3.10 Zener diodes . . . . . . . . . . . . . . . . . . .
3.11 Special-purpose diodes . . . . . . . . . . . . .
3.11.1 Schottky diodes . . . . . . . . . . . . . . . .
3.11.2 Tunnel diodes . . . . . . . . . . . . . . . . .
3.11.3 Light-emitting diodes . . . . . . . . . . . . .
3.11.4 Laser diodes . . . . . . . . . . . . . . . . . .
3.11.5 Photodiodes . . . . . . . . . . . . . . . . . .
3.11.6 Solar cells . . . . . . . . . . . . . . . . . . .
3.11.7 Varicap or varactor diodes . . . . . . . . . .
3.11.8 Snap diode . . . . . . . . . . . . . . . . . . .
3.11.9 PIN diodes . . . . . . . . . . . . . . . . . . .
3.11.10 IMPATT diode . . . . . . . . . . . . . . . .
3.11.11 Gunn diode . . . . . . . . . . . . . . . . . .
3.11.12 Shockley diode . . . . . . . . . . . . . . . . .
3.11.13 Constant-current diodes . . . . . . . . . . . .
3.12 Other diode technologies . . . . . . . . . . .

93

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . .
. . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . .

99
103
104
111
112
116
118
124
127
127
128
129
137
137
138
140
145
146
148
153
153
154
154
155
156
156
157

94

CHAPTER 3. DIODES AND RECTIFIERS


3.12.1 SiC diodes . .
3.12.2 Polymer diode
3.13 SPICE models . .
Bibliography . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . .
. . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

. . . . . . .
. . . . . . .
. . . . . .
. . . . . .

157
157
157
165

A diode is an electrical device allowing current to move through it in one direction with far
greater ease than in the other. The most common kind of diode in modern circuit design is the
semiconductor diode, although other diode technologies exist. Semiconductor diodes are symbolized
in schematic diagrams such as Figure 3.1. The term diode is customarily reserved for small signal
devices, I 1 A. The term rectifier is used for power devices, I > 1 A.

Figure 3.1: Semiconductor diode schematic symbol: Arrows indicate the direction of electron current
flow.
When placed in a simple battery-lamp circuit, the diode will either allow or prevent current
through the lamp, depending on the polarity of the applied voltage. (Figure 3.2)

+
(a)

(b)

Figure 3.2: Diode operation: (a) Current flow is permitted; the diode is forward biased. (b) Current
flow is prohibited; the diode is reversed biased.
When the polarity of the battery is such that electrons are allowed to flow through the diode,
the diode is said to be forward-biased. Conversely, when the battery is backward and the diode
blocks current, the diode is said to be reverse-biased. A diode may be thought of as like a switch:
closed when forward-biased and open when reverse-biased.
Oddly enough, the direction of the diode symbols arrowhead points against the direction of
electron flow. This is because the diode symbol was invented by engineers, who predominantly use
conventional flow notation in their schematics, showing current as a flow of charge from the positive
(+) side of the voltage source to the negative (-). This convention holds true for all semiconductor
symbols possessing arrowheads: the arrow points in the permitted direction of conventional flow,
and against the permitted direction of electron flow.
Diode behavior is analogous to the behavior of a hydraulic device called a check valve. A check
valve allows fluid flow through it in only one direction as in Figure 3.3.
Check valves are essentially pressure-operated devices: they open and allow flow if the pressure
across them is of the correct polarity to open the gate (in the analogy shown, greater fluid pressure
on the right than on the left). If the pressure is of the opposite polarity, the pressure difference
across the check valve will close and hold the gate so that no flow occurs.

95

Hydraulic
check valve

(a)

Flow permitted

(b)

Flow prohibited

Figure 3.3: Hydraulic check valve analogy: (a) Electron current flow permitted. (b) Current flow
prohibited.
Like check valves, diodes are essentially pressure- operated (voltage-operated) devices. The
essential difference between forward-bias and reverse-bias is the polarity of the voltage dropped
across the diode. Lets take a closer look at the simple battery-diode-lamp circuit shown earlier, this
time investigating voltage drops across the various components in Figure 3.4.
6.0 V

0.7 V

V
V

A COM

A COM

+ -

- +

5.3 V

6V
(a)

A COM

0.0 V

6V
(b)

A COM

Figure 3.4: Diode circuit voltage measurements: (a) Forward biased. (b) Reverse biased.
A forward-biased diode conducts current and drops a small voltage across it, leaving most of the
battery voltage dropped across the lamp. If the batterys polarity is reversed, the diode becomes
reverse-biased, and drops all of the batterys voltage leaving none for the lamp. If we consider the
diode to be a self-actuating switch (closed in the forward-bias mode and open in the reverse-bias
mode), this behavior makes sense. The most substantial difference is that the diode drops a lot more
voltage when conducting than the average mechanical switch (0.7 volts versus tens of millivolts).
This forward-bias voltage drop exhibited by the diode is due to the action of the depletion region
formed by the P-N junction under the influence of an applied voltage. If no voltage applied is
across a semiconductor diode, a thin depletion region exists around the region of the P-N junction,
preventing current flow. (Figure 3.5 (a)) The depletion region is almost devoid of available charge
carriers, and acts as an insulator:
The schematic symbol of the diode is shown in Figure 3.5 (b) such that the anode (pointing
end) corresponds to the P-type semiconductor at (a). The cathode bar, non-pointing end, at (b)
corresponds to the N-type material at (a). Also note that the cathode stripe on the physical part

96

CHAPTER 3. DIODES AND RECTIFIERS

P-type
material

N-type
material

P-N junction representation

(a)
Depletion region
Anode

Cathode
Schematic symbol

(b)

Stripe marks cathode

Real component appearance


(c)

Figure 3.5: Diode representations: PN-junction model, schematic symbol, physical part.
(c) corresponds to the cathode on the symbol.
If a reverse-biasing voltage is applied across the P-N junction, this depletion region expands,
further resisting any current through it. (Figure 3.6)

P
Reverse-biased

N
Depletion region

Figure 3.6: Depletion region expands with reverse bias.


Conversely, if a forward-biasing voltage is applied across the P-N junction, the depletion region
collapses becoming thinner. The diode becomes less resistive to current through it. In order for a
sustained current to go through the diode; though, the depletion region must be fully collapsed by
the applied voltage. This takes a certain minimum voltage to accomplish, called the forward voltage
as illustrated in Figure 3.7.
For silicon diodes, the typical forward voltage is 0.7 volts, nominal. For germanium diodes, the
forward voltage is only 0.3 volts. The chemical constituency of the P-N junction comprising the
diode accounts for its nominal forward voltage figure, which is why silicon and germanium diodes
have such different forward voltages. Forward voltage drop remains approximately constant for a
wide range of diode currents, meaning that diode voltage drop is not like that of a resistor or even

97

0.4 V

0.7 V

Partial forward-biased

P
(a)

Forward-biased

N
Depletion region

P
(b)

Depletion region fully collapsed

Figure 3.7: Inceasing forward bias from (a) to (b) decreases depletion region thickness.

a normal (closed) switch. For most simplified circuit analysis, the voltage drop across a conducting
diode may be considered constant at the nominal figure and not related to the amount of current.

Actually, forward voltage drop is more complex. An equation describes the exact current through
a diode, given the voltage dropped across the junction, the temperature of the junction, and several
physical constants. It is commonly known as the diode equation:

ID = IS (eqVD/NkT - 1)
Where,
ID = Diode current in amps
IS = Saturation current in amps
(typically 1 x 10-12 amps)
e = Eulers constant (~ 2.718281828)
q = charge of electron (1.6 x 10-19 coulombs)
VD = Voltage applied across diode in volts
N = "Nonideality" or "emission" coefficient
(typically between 1 and 2)
k = Boltzmanns constant (1.38 x 10-23)
T = Junction temperature in Kelvins
The term kT/q describes the voltage produced within the P-N junction due to the action of
temperature, and is called the thermal voltage, or Vt of the junction. At room temperature, this is
about 26 millivolts. Knowing this, and assuming a nonideality coefficient of 1, we may simplify
the diode equation and re-write it as such:

98

CHAPTER 3. DIODES AND RECTIFIERS

ID = IS (eVD/0.026 -1)
Where,
ID = Diode current in amps
IS = Saturation current in amps
(typically 1 x 10-12 amps)
e = Eulers constant (~ 2.718281828)
VD = Voltage applied across diode in volts
You need not be familiar with the diode equation to analyze simple diode circuits. Just
understand that the voltage dropped across a current-conducting diode does change with the amount
of current going through it, but that this change is fairly small over a wide range of currents. This is
why many textbooks simply say the voltage drop across a conducting, semiconductor diode remains
constant at 0.7 volts for silicon and 0.3 volts for germanium. However, some circuits intentionally
make use of the P-N junctions inherent exponential current/voltage relationship and thus can only
be understood in the context of this equation. Also, since temperature is a factor in the diode
equation, a forward-biased P-N junction may also be used as a temperature-sensing device, and thus
can only be understood if one has a conceptual grasp on this mathematical relationship.
A reverse-biased diode prevents current from going through it, due to the expanded depletion
region. In actuality, a very small amount of current can and does go through a reverse-biased
diode, called the leakage current, but it can be ignored for most purposes. The ability of a diode
to withstand reverse-bias voltages is limited, as it is for any insulator. If the applied reverse-bias
voltage becomes too great, the diode will experience a condition known as breakdown (Figure 3.8),
which is usually destructive. A diodes maximum reverse-bias voltage rating is known as the Peak
Inverse Voltage, or PIV, and may be obtained from the manufacturer. Like forward voltage, the PIV
rating of a diode varies with temperature, except that PIV increases with increased temperature
and decreases as the diode becomes cooler exactly opposite that of forward voltage.
Typically, the PIV rating of a generic rectifier diode is at least 50 volts at room temperature.
Diodes with PIV ratings in the many thousands of volts are available for modest prices.
REVIEW:
A diode is an electrical component acting as a one-way valve for current.
When voltage is applied across a diode in such a way that the diode allows current, the diode
is said to be forward-biased.
When voltage is applied across a diode in such a way that the diode prohibits current, the
diode is said to be reverse-biased.
The voltage dropped across a conducting, forward-biased diode is called the forward voltage.
Forward voltage for a diode varies only slightly for changes in forward current and temperature,
and is fixed by the chemical composition of the P-N junction.
Silicon diodes have a forward voltage of approximately 0.7 volts.
Germanium diodes have a forward voltage of approximately 0.3 volts.

3.1.

METER CHECK OF A DIODE

99

ID
forward

reverse-bias

forward-bias

0.7
V

breakdown!

VD

reverse

Figure 3.8: Diode curve: showing knee at 0.7 V forward bias for Si, and reverse breakdown.
The maximum reverse-bias voltage that a diode can withstand without breaking down is
called the Peak Inverse Voltage, or PIV rating.

3.1

Meter check of a diode

Being able to determine the polarity (cathode versus anode) and basic functionality of a diode is a
very important skill for the electronics hobbyist or technician to have. Since we know that a diode
is essentially nothing more than a one-way valve for electricity, it makes sense we should be able to
verify its one-way nature using a DC (battery-powered) ohmmeter as in Figure 3.9. Connected one
way across the diode, the meter should show a very low resistance at (a). Connected the other way
across the diode, it should show a very high resistance at (b) (OL on some digital meter models).
Of course, to determine which end of the diode is the cathode and which is the anode, you must
know with certainty which test lead of the meter is positive (+) and which is negative (-) when set
to the resistance or function. With most digital multimeters Ive seen, the red lead becomes
positive and the black lead negative when set to measure resistance, in accordance with standard
electronics color-code convention. However, this is not guaranteed for all meters. Many analog
multimeters, for example, actually make their black leads positive (+) and their red leads negative
(-) when switched to the resistance function, because it is easier to manufacture it that way!
One problem with using an ohmmeter to check a diode is that the readings obtained only have
qualitative value, not quantitative. In other words, an ohmmeter only tells you which way the diode
conducts; the low-value resistance indication obtained while conducting is useless. If an ohmmeter
shows a value of 1.73 ohms while forward-biasing a diode, that figure of 1.73 doesnt represent
any real-world quantity useful to us as technicians or circuit designers. It neither represents the
forward voltage drop nor any bulk resistance in the semiconductor material of the diode itself,
but rather is a figure dependent upon both quantities and will vary substantially with the particular

100

CHAPTER 3. DIODES AND RECTIFIERS

+
Anode

OFF

+
Cathode

OFF

Cathode
A

Anode

COM

COM

(a)

(b)

Figure 3.9: Determination of diode polarity: (a) Low resistance indicates forward bias, black lead is
cathode and red lead anode (for most meters) (b) Reversing leads shows high resistance indicating
reverse bias.
ohmmeter used to take the reading.
For this reason, some digital multimeter manufacturers equip their meters with a special diode
check function which displays the actual forward voltage drop of the diode in volts, rather than a
resistance figure in ohms. These meters work by forcing a small current through the diode and
measuring the voltage dropped between the two test leads. (Figure 3.10)

OFF

Anode

COM

Cathode

Figure 3.10: Meter with a Diode check function displays the forward voltage drop of 0.548 volts
instead of a low resistance.
The forward voltage reading obtained with such a meter will typically be less than the normal
drop of 0.7 volts for silicon and 0.3 volts for germanium, because the current provided by the meter
is of trivial proportions. If a multimeter with diode-check function isnt available, or you would like
to measure a diodes forward voltage drop at some non-trivial current, the circuit of Figure 3.11 may
be constructed using a battery, resistor, and voltmeter
Connecting the diode backwards to this testing circuit will simply result in the voltmeter indi-

3.1.

METER CHECK OF A DIODE

101

A
OFF

Resistor

Diode

V
-

COM

Battery

(a)

(b)

Figure 3.11: Measuring forward voltage of a diode withoutdiode check meter function: (a)
Schematic diagram. (b) Pictorial diagram.
cating the full voltage of the battery.
If this circuit were designed to provide a constant or nearly constant current through the diode
despite changes in forward voltage drop, it could be used as the basis of a temperature-measurement
instrument, the voltage measured across the diode being inversely proportional to diode junction
temperature. Of course, diode current should be kept to a minimum to avoid self-heating (the diode
dissipating substantial amounts of heat energy), which would interfere with temperature measurement.
Beware that some digital multimeters equipped with a diode check function may output a very
low test voltage (less than 0.3 volts) when set to the regular resistance () function: too low to
fully collapse the depletion region of a PN junction. The philosophy here is that the diode check
function is to be used for testing semiconductor devices, and the resistance function for anything
else. By using a very low test voltage to measure resistance, it is easier for a technician to measure
the resistance of non-semiconductor components connected to semiconductor components, since the
semiconductor component junctions will not become forward-biased with such low voltages.
Consider the example of a resistor and diode connected in parallel, soldered in place on a printed
circuit board (PCB). Normally, one would have to unsolder the resistor from the circuit (disconnect
it from all other components) before measuring its resistance, otherwise any parallel-connected
components would affect the reading obtained. When using a multimeter which outputs a very low
test voltage to the probes in the resistance function mode, the diodes PN junction will not have
enough voltage impressed across it to become forward-biased, and will only pass negligible current.
Consequently, the meter sees the diode as an open (no continuity), and only registers the resistors
resistance. (Figure 3.12)
If such an ohmmeter were used to test a diode, it would indicate a very high resistance (many
mega-ohms) even if connected to the diode in the correct (forward-biased) direction. (Figure 3.13)
Reverse voltage strength of a diode is not as easily tested, because exceeding a normal diodes
PIV usually results in destruction of the diode. Special types of diodes, though, which are designed
to break down in reverse-bias mode without damage (called zener diodes), which are tested with

102

CHAPTER 3. DIODES AND RECTIFIERS

OFF

D1

R1

1 k

Printed circuit board

COM

Figure 3.12: Ohmmeter equipped with a low test voltage (<0.7 V) does not see diodes allowing it
to measure parallel resistors.

OFF

COM

Figure 3.13: Ohmmeter equipped with a low test voltage, too low to forward bias diodes, does not
see diodes.

3.2.

DIODE RATINGS

103

the same voltage source / resistor / voltmeter circuit, provided that the voltage source is of high
enough value to force the diode into its breakdown region. More on this subject in a later section of
this chapter.
REVIEW:
An ohmmeter may be used to qualitatively check diode function. There should be low resistance
measured one way and very high resistance measured the other way. When using an ohmmeter
for this purpose, be sure you know which test lead is positive and which is negative! The actual
polarity may not follow the colors of the leads as you might expect, depending on the particular
design of meter.
Some multimeters provide a diode check function that displays the actual forward voltage of
the diode when its conducting current. Such meters typically indicate a slightly lower forward
voltage than what is nominal for a diode, due to the very small amount of current used
during the check.

3.2

Diode ratings

In addition to forward voltage drop (Vf ) and peak inverse voltage (PIV), there are many other ratings
of diodes important to circuit design and component selection. Semiconductor manufacturers provide
detailed specifications on their products diodes included in publications known as datasheets.
Datasheets for a wide variety of semiconductor components may be found in reference books and
on the internet. I prefer the internet as a source of component specifications because all the data
obtained from manufacturer websites are up-to-date.
A typical diode datasheet will contain figures for the following parameters:
Maximum repetitive reverse voltage = VRRM , the maximum amount of voltage the diode can
withstand in reverse-bias mode, in repeated pulses. Ideally, this figure would be infinite.
Maximum DC reverse voltage = VR or VDC , the maximum amount of voltage the diode can
withstand in reverse-bias mode on a continual basis. Ideally, this figure would be infinite.
Maximum forward voltage = VF , usually specified at the diodes rated forward current. Ideally,
this figure would be zero: the diode providing no opposition whatsoever to forward current. In
reality, the forward voltage is described by the diode equation.
Maximum (average) forward current = IF (AV ) , the maximum average amount of current the
diode is able to conduct in forward bias mode. This is fundamentally a thermal limitation: how
much heat can the PN junction handle, given that dissipation power is equal to current (I) multiplied
by voltage (V or E) and forward voltage is dependent upon both current and junction temperature.
Ideally, this figure would be infinite.
Maximum (peak or surge) forward current = IF SM or if (surge) , the maximum peak amount of
current the diode is able to conduct in forward bias mode. Again, this rating is limited by the
diode junctions thermal capacity, and is usually much higher than the average current rating due
to thermal inertia (the fact that it takes a finite amount of time for the diode to reach maximum
temperature for a given current). Ideally, this figure would be infinite.
Maximum total dissipation = PD , the amount of power (in watts) allowable for the diode to
dissipate, given the dissipation (P=IE) of diode current multiplied by diode voltage drop, and also

104

CHAPTER 3. DIODES AND RECTIFIERS

the dissipation (P=I2 R) of diode current squared multiplied by bulk resistance. Fundamentally
limited by the diodes thermal capacity (ability to tolerate high temperatures).
Operating junction temperature = TJ , the maximum allowable temperature for the diodes PN
junction, usually given in degrees Celsius (o C). Heat is the Achilles heel of semiconductor devices:
they must be kept cool to function properly and give long service life.
Storage temperature range = TST G , the range of allowable temperatures for storing a diode (unpowered). Sometimes given in conjunction with operating junction temperature (T J ), because the
maximum storage temperature and the maximum operating temperature ratings are often identical. If anything, though, maximum storage temperature rating will be greater than the maximum
operating temperature rating.
Thermal resistance = R(), the temperature difference between junction and outside air (R() JA )
or between junction and leads (R()JL ) for a given power dissipation. Expressed in units of degrees
Celsius per watt (o C/W). Ideally, this figure would be zero, meaning that the diode package was
a perfect thermal conductor and radiator, able to transfer all heat energy from the junction to the
outside air (or to the leads) with no difference in temperature across the thickness of the diode
package. A high thermal resistance means that the diode will build up excessive temperature at the
junction (where its critical) despite best efforts at cooling the outside of the diode, and thus will
limit its maximum power dissipation.
Maximum reverse current = IR , the amount of current through the diode in reverse-bias operation, with the maximum rated inverse voltage applied (VDC ). Sometimes referred to as leakage
current. Ideally, this figure would be zero, as a perfect diode would block all current when reversebiased. In reality, it is very small compared to the maximum forward current.
Typical junction capacitance = CJ , the typical amount of capacitance intrinsic to the junction,
due to the depletion region acting as a dielectric separating the anode and cathode connections.
This is usually a very small figure, measured in the range of picofarads (pF).
Reverse recovery time = trr , the amount of time it takes for a diode to turn off when the
voltage across it alternates from forward-bias to reverse-bias polarity. Ideally, this figure would be
zero: the diode halting conduction immediately upon polarity reversal. For a typical rectifier diode,
reverse recovery time is in the range of tens of microseconds; for a fast switching diode, it may
only be a few nanoseconds.
Most of these parameters vary with temperature or other operating conditions, and so a single
figure fails to fully describe any given rating. Therefore, manufacturers provide graphs of component
ratings plotted against other variables (such as temperature), so that the circuit designer has a better
idea of what the device is capable of.

3.3

Rectifier circuits

Now we come to the most popular application of the diode: rectification. Simply defined, rectification is the conversion of alternating current (AC) to direct current (DC). This involves a device
that only allows one-way flow of electrons. As we have seen, this is exactly what a semiconductor
diode does. The simplest kind of rectifier circuit is the half-wave rectifier. It only allows one half of
an AC waveform to pass through to the load. (Figure 3.14)
For most power applications, half-wave rectification is insufficient for the task. The harmonic
content of the rectifiers output waveform is very large and consequently difficult to filter. Further-

3.3.

RECTIFIER CIRCUITS

105

AC
voltage
source

Load

Figure 3.14: Half-wave rectifier circuit.


more, the AC power source only supplies power to the load once every half-cycle, meaning that much
of its capacity is unused. Half-wave rectification is, however, a very simple way to reduce power to
a resistive load. Some two-position lamp dimmer switches apply full AC power to the lamp filament
for full brightness and then half-wave rectify it for a lesser light output. (Figure 3.15)
Bright
AC
voltage
source

Dim

Figure 3.15: Half-wave rectifier application: Two level lamp dimmer.


In the Dim switch position, the incandescent lamp receives approximately one-half the power
it would normally receive operating on full-wave AC. Because the half-wave rectified power pulses far
more rapidly than the filament has time to heat up and cool down, the lamp does not blink. Instead,
its filament merely operates at a lesser temperature than normal, providing less light output. This
principle of pulsing power rapidly to a slow-responding load device to control the electrical power
sent to it is common in the world of industrial electronics. Since the controlling device (the diode,
in this case) is either fully conducting or fully nonconducting at any given time, it dissipates little
heat energy while controlling load power, making this method of power control very energy-efficient.
This circuit is perhaps the crudest possible method of pulsing power to a load, but it suffices as a
proof-of-concept application.
If we need to rectify AC power to obtain the full use of both half-cycles of the sine wave, a
different rectifier circuit configuration must be used. Such a circuit is called a full-wave rectifier.
One kind of full-wave rectifier, called the center-tap design, uses a transformer with a center-tapped
secondary winding and two diodes, as in Figure 3.16.
This circuits operation is easily understood one half-cycle at a time. Consider the first half-cycle,
when the source voltage polarity is positive (+) on top and negative (-) on bottom. At this time,
only the top diode is conducting; the bottom diode is blocking current, and the load sees the first
half of the sine wave, positive on top and negative on bottom. Only the top half of the transformers
secondary winding carries current during this half-cycle as in Figure 3.17.
During the next half-cycle, the AC polarity reverses. Now, the other diode and the other half of
the transformers secondary winding carry current while the portions of the circuit formerly carrying
current during the last half-cycle sit idle. The load still sees half of a sine wave, of the same polarity

106

CHAPTER 3. DIODES AND RECTIFIERS

AC
voltage
source

+
Load

Figure 3.16: Full-wave rectifier, center-tapped design.

+
-

Figure 3.17: Full-wave center-tap rectifier: Top half of secondary conducts during positive half-cycle
of input, delivering positive half-cycle to load..
as before: positive on top and negative on bottom. (Figure 3.18)

+
-

Figure 3.18: Full-wave center-tap rectifier: During negative input half-cycle, bottom half of secondary conducts, delivering a positive half-cycle to the load.
One disadvantage of this full-wave rectifier design is the necessity of a transformer with a centertapped secondary winding. If the circuit in question is one of high power, the size and expense of
a suitable transformer is significant. Consequently, the center-tap rectifier design is only seen in
low-power applications.
The full-wave center-tapped rectifier polarity at the load may be reversed by changing the direction of the diodes. Furthermore, the reversed diodes can be paralleled with an existing positiveoutput rectifier. The result is dual-polarity full-wave center-tapped rectifier in Figure 3.19. Note
that the connectivity of the diodes themselves is the same configuration as a bridge.
Another, more popular full-wave rectifier design exists, and it is built around a four-diode bridge

3.3.

RECTIFIER CIRCUITS

107
Loads

+
+
-

AC voltage source

Figure 3.19: Dual polarity full-wave center tap rectifier


configuration. For obvious reasons, this design is called a full-wave bridge. (Figure 3.20)

AC
voltage
source

+
Load

Figure 3.20: Full-wave bridge rectifier.


Current directions for the full-wave bridge rectifier circuit are as shown in Figure 3.21 for positive
half-cycle and Figure 3.22 for negative half-cycles of the AC source waveform. Note that regardless
of the polarity of the input, the current flows in the same direction through the load. That is, the
negative half-cycle of source is a positive half-cycle at the load. The current flow is through two
diodes in series for both polarities. Thus, two diode drops of the source voltage are lost (0.72=1.4
V for Si) in the diodes. This is a disadvantage compared with a full-wave center-tap design. This
disadvantage is only a problem in very low voltage power supplies.

+
-

+
-

Figure 3.21: Full-wave bridge rectifier: Electron flow for positive half-cycles.

108

CHAPTER 3. DIODES AND RECTIFIERS

+
-

Figure 3.22: Full-wave bridge rectifier: Electron flow for negative half=cycles.
Remembering the proper layout of diodes in a full-wave bridge rectifier circuit can often be
frustrating to the new student of electronics. Ive found that an alternative representation of this
circuit is easier both to remember and to comprehend. Its the exact same circuit, except all diodes
are drawn in a horizontal attitude, all pointing the same direction. (Figure 3.23)

AC
voltage
source

+
Load

Figure 3.23: Alternative layout style for Full-wave bridge rectifier.


One advantage of remembering this layout for a bridge rectifier circuit is that it expands easily
into a polyphase version in Figure 3.24.

3-phase
AC source

+
Load

Figure 3.24: Three-phase full-wave bridge rectifier circuit.


Each three-phase line connects between a pair of diodes: one to route power to the positive

3.3.

RECTIFIER CIRCUITS

109

(+) side of the load, and the other to route power to the negative (-) side of the load. Polyphase
systems with more than three phases are easily accommodated into a bridge rectifier scheme. Take
for instance the six-phase bridge rectifier circuit in Figure 3.25.

6-phase
AC source

+
Load

Figure 3.25: Six-phase full-wave bridge rectifier circuit.


When polyphase AC is rectified, the phase-shifted pulses overlap each other to produce a DC
output that is much smoother (has less AC content) than that produced by the rectification of
single-phase AC. This is a decided advantage in high-power rectifier circuits, where the sheer physical
size of filtering components would be prohibitive but low-noise DC power must be obtained. The
diagram in Figure 3.26 shows the full-wave rectification of three-phase AC.
1

TIME
Resultant DC waveform

Figure 3.26: Three-phase AC and 3-phase full-wave rectifier output.


In any case of rectification single-phase or polyphase the amount of AC voltage mixed with
the rectifiers DC output is called ripple voltage. In most cases, since pure DC is the desired
goal, ripple voltage is undesirable. If the power levels are not too great, filtering networks may be
employed to reduce the amount of ripple in the output voltage.

110

CHAPTER 3. DIODES AND RECTIFIERS

Sometimes, the method of rectification is referred to by counting the number of DC pulses


output for every 360o of electrical rotation. A single-phase, half-wave rectifier circuit, then, would
be called a 1-pulse rectifier, because it produces a single pulse during the time of one complete cycle
(360o ) of the AC waveform. A single-phase, full-wave rectifier (regardless of design, center-tap or
bridge) would be called a 2-pulse rectifier, because it outputs two pulses of DC during one AC cycles
worth of time. A three-phase full-wave rectifier would be called a 6-pulse unit.
Modern electrical engineering convention further describes the function of a rectifier circuit by
using a three-field notation of phases, ways, and number of pulses. A single-phase, half-wave rectifier
circuit is given the somewhat cryptic designation of 1Ph1W1P (1 phase, 1 way, 1 pulse), meaning
that the AC supply voltage is single-phase, that current on each phase of the AC supply lines
moves in only one direction (way), and that there is a single pulse of DC produced for every 360 o
of electrical rotation. A single-phase, full-wave, center-tap rectifier circuit would be designated as
1Ph1W2P in this notational system: 1 phase, 1 way or direction of current in each winding half, and
2 pulses or output voltage per cycle. A single-phase, full-wave, bridge rectifier would be designated
as 1Ph2W2P: the same as for the center-tap design, except current can go both ways through the
AC lines instead of just one way. The three-phase bridge rectifier circuit shown earlier would be
called a 3Ph2W6P rectifier.
Is it possible to obtain more pulses than twice the number of phases in a rectifier circuit? The
answer to this question is yes: especially in polyphase circuits. Through the creative use of transformers, sets of full-wave rectifiers may be paralleled in such a way that more than six pulses of DC
are produced for three phases of AC. A 30o phase shift is introduced from primary to secondary of a
three-phase transformer when the winding configurations are not of the same type. In other words,
a transformer connected either Y- or -Y will exhibit this 30o phase shift, while a transformer
connected Y-Y or - will not. This phenomenon may be exploited by having one transformer
connected Y-Y feed a bridge rectifier, and have another transformer connected Y- feed a second
bridge rectifier, then parallel the DC outputs of both rectifiers. (Figure 3.27) Since the ripple voltage
waveforms of the two rectifiers outputs are phase-shifted 30o from one another, their superposition
results in less ripple than either rectifier output considered separately: 12 pulses per 360 o instead of
just six:
REVIEW:
Rectification is the conversion of alternating current (AC) to direct current (DC).
A half-wave rectifier is a circuit that allows only one half-cycle of the AC voltage waveform to
be applied to the load, resulting in one non-alternating polarity across it. The resulting DC
delivered to the load pulsates significantly.
A full-wave rectifier is a circuit that converts both half-cycles of the AC voltage waveform to
an unbroken series of voltage pulses of the same polarity. The resulting DC delivered to the
load doesnt pulsate as much.
Polyphase alternating current, when rectified, gives a much smoother DC waveform (less
ripple voltage) than rectified single-phase AC.

3.4.

PEAK DETECTOR

111

3Ph2W12P rectifier circuit


Secondary

Primary
3-phase
AC input

DC
output

Secondary

Figure 3.27: Polyphase rectifier circuit: 3-phase 2-way 12-pulse (3Ph2W12P)

3.4

Peak detector

A peak detector is a series connection of a diode and a capacitor outputting a DC voltage equal to
the peak value of the applied AC signal. The circuit is shown in Figure 3.28 with the corresponding
SPICE net list. An AC voltage source applied to the peak detector, charges the capacitor to the
peak of the input. The diode conducts positive half cycles, charging the capacitor to the waveform
peak. When the input waveform falls below the DC peak stored on the capacitor, the diode is
reverse biased, blocking current flow from capacitor back to the source. Thus, the capacitor retains
the peak value even as the waveform drops to zero. Another view of the peak detector is that it is
the same as a half-wave rectifier with a filter capacitor added to the output.

1.0 k
5 Vp-p
0 Voffset
1 kHz 0

2
V(2)
output

0.1 uF

*SPICE 03441.eps
C1 2 0 0.1u
R1 1 3 1.0k
V1 1 0 SIN(0 5 1k)
D1 3 2 diode
.model diode d
.tran 0.01m 50mm
.end

Figure 3.28: Peak detector: Diode conducts on positive half cycles charging capacitor to the peak
voltage (less diode forward drop).
It takes a few cycles for the capacitor to charge to the peak as in Figure 3.29 due to the series
resistance (RC time constant). Why does the capacitor not charge all the way to 5 V? It would
charge to 5 V if an ideal diode were obtainable. However, the silicon diode has a forward voltage
drop of 0.7 V which subtracts from the 5 V peak of the input.

112

CHAPTER 3. DIODES AND RECTIFIERS

Figure 3.29: Peak detector: Capacitor charges to peak within a few cycles.
The circuit in Figure 3.28 could represent a DC power supply based on a half-wave rectifier. The
resistance would be a few Ohms instead of 1 k due to a transformer secondary winding replacing
the voltage source and resistor. A larger filter capacitor would be used. A power supply based
on a 60 Hz source with a filter of a few hundred F could supply up to 100 mA. Half-wave supplies
seldom supply more due to the difficulty of filtering a half-wave.
The peak detector may be combined with other components to build a crystal radio ( page 376).

3.5

Clipper circuits

A circuit which removes the peak of a waveform is known as a clipper. A negative clipper is shown
in Figure 3.30. This schematic diagram was produced with Xcircuit schematic capture program.
Xcircuit produced the SPICE net list Figure 3.30, except for the second, and next to last pair of
lines which were inserted with a text editor.

2
1.0 k

5 Vp
0 Voffset
0 1 kHz

V(2)
output

*SPICE 03437.eps
* A K ModelName
D1 0 2 diode
R1 2 1 1.0k
V1 1 0 SIN(0 5 1k)
.model diode d
.tran .05m 3m
.end

Figure 3.30: Clipper: clips negative peak at -0.7 V.

3.5.

CLIPPER CIRCUITS

113

During the positive half cycle of the 5 V peak input, the diode is reversed biased. The diode does
not conduct. It is as if the diode were not there. The positive half cycle is unchanged at the output
V(2) in Figure 3.31. Since the output positive peaks actually overlays the input sinewave V(1), the
input has been shifted upward in the plot for clarity. In Nutmeg, the SPICE display module, the
command plot v(1)+1) accomplishes this.

Figure 3.31: V(1)+1 is actually V(1), a 5 Vptp sinewave, offset by 1 V for display clarity. V(2)
output is clipped at -0.7 V, by diode D1.
During the negative half cycle of sinewave input of Figure 3.31, the diode is forward biased, that
is, conducting. The negative half cycle of the sinewave is shorted out. The negative half cycle of
V(2) would be clipped at 0 V for an ideal diode. The waveform is clipped at -0.7 V due to the
forward voltage drop of the silicon diode. The spice model defaults to 0.7 V unless parameters in
the model statement specify otherwise. Germanium or Schottky diodes clip at lower voltages.
Closer examination of the negative clipped peak (Figure 3.31) reveals that it follows the input
for a slight period of time while the sinewave is moving toward -0.7 V. The clipping action is only
effective after the input sinewave exceeds -0.7 V. The diode is not conducting for the complete half
cycle, though, during most of it.
The addition of an anti-parallel diode to the existing diode in Figure 3.30 yields the symmetrical
clipper in Figure 3.32.
Diode D1 clips the negative peak at -0.7 V as before. The additional diode D2 conducts for
positive half cycles of the sine wave as it exceeds 0.7 V, the forward diode drop. The remainder of
the voltage drops across the series resistor. Thus, both peaks of the input sinewave are clipped in
Figure 3.33. The net list is in Figure 3.32
The most general form of the diode clipper is shown in Figure 3.34. For an ideal diode, the
clipping occurs at the level of the clipping voltage, V1 and V2. However, the voltage sources have
been adjusted to account for the 0.7 V forward drop of the real silicon diodes. D1 clips at 1.3V
+0.7V=2.0V when the diode begins to conduct. D2 clips at -2.3V -0.7V=-3.0V when D2 conducts.

114

CHAPTER 3. DIODES AND RECTIFIERS

2
1.0 k

5 Vp
0 Voffset
0 1 kHz

D2

D1

*SPICE 03438.eps
D1 0 2 diode
D2 2 0 diode
R1 2 1 1.0k
V1 1 0 SIN(0 5 1k)
.model diode d
.tran 0.05m 3m
.end

Figure 3.32: Symmetrical clipper: Anti-parallel diodes clip both positive and negative peak, leaving
a 0.7 V output.

Figure 3.33: Diode D1 clips at -0.7 V as it conducts during negative peaks. D2 conducts for positive
peaks, clipping at 0.7V.

1.0 k

V3
5 Vp
0 Voffset
1 kHz

D1
3
V1
+
1.3
V

V2
D2
4
V2
+
-2.3
V

*SPICE 03439.eps
V1 3 0 1.3
V2 4 0 -2.3
D1 2 3 diode
D2 4 2 diode
R1 2 1 1.0k
V3 1 0 SIN(0 5 1k)
.model diode d
.tran 0.05m 3m
.end

Figure 3.34: D1 clips the input sinewave at 2V. D2 clips at -3V.

3.5.

CLIPPER CIRCUITS

115

The clipper in Figure 3.34 does not have to clip both levels. To clip at one level with one diode
and one voltage source, remove the other diode and source.
The net list is in Figure 3.34. The waveforms in Figure 3.35 show the clipping of v(1) at output
v(2).

Figure 3.35: D1 clips the sinewave at 2V. D2 clips at -3V.


There is also a zener diode clipper circuit in the Zener diode section. A zener diode replaces
both the diode and the DC voltage source.
A practical application of a clipper is to prevent an amplified speech signal from overdriving a
radio transmitter in Figure 3.36. Over driving the transmitter generates spurious radio signals which
causes interference with other stations. The clipper is a protective measure.

clipper
preamp

transimtter

microphone

Figure 3.36: Clipper prevents over driving radio transmitter by voice peaks.
A sinewave may be squared up by overdriving a clipper. Another clipper application is the
protection of exposed inputs of integrated circuits. The input of the IC is connected to a pair of
diodes as at node 2 of Figure 3.34. The voltage sources are replaced by the power supply rails of
the IC. For example, CMOS ICs use 0V and +5 V. Analog amplifiers might use 12V for the V1

116

CHAPTER 3. DIODES AND RECTIFIERS

and V2 sources.
REVIEW
A resistor and diode driven by an AC voltage source clips the signal observed across the diode.
A pair of anti-parallel Si diodes clip symmetrically at 0.7V
The grounded end of a clipper diode(s) can be disconnected and wired to a DC voltage to clip
at an arbitrary level.
A clipper can serve as a protective measure, preventing a signal from exceeding the clip limits.

3.6

Clamper circuits

The circuits in Figure 3.37 are known as clampers or DC restorers. The corresponding netlist is
in Figure 3.38. These circuits clamp a peak of a waveform to a specific DC level compared with
a capacitively coupled signal which swings about its average DC level (usually 0V). If the diode is
removed from the clamper, it defaults to a simple coupling capacitor no clamping.
What is the clamp voltage? And, which peak gets clamped? In Figure 3.37 (a) the clamp voltage
is 0 V ignoring diode drop, (more exactly 0.7 V with Si diode drop). In Figure 3.38, the positive
peak of V(1) is clamped to the 0 V (0.7 V) clamp level. Why is this? On the first positive half cycle,
the diode conducts charging the capacitor left end to +5 V (4.3 V). This is -5 V (-4.3 V) on the right
end at V(1,4). Note the polarity marked on the capacitor in Figure 3.37 (a). The right end of the
capacitor is -5 V DC (-4.3 V) with respect to ground. It also has an AC 5 V peak sinewave coupled
across it from source V(4) to node 1. The sum of the two is a 5 V peak sine riding on a - 5 V DC
(-4.3 V) level. The diode only conducts on successive positive excursions of source V(4) if the peak
V(4) exceeds the charge on the capacitor. This only happens if the charge on the capacitor drained
off due to a load, not shown. The charge on the capacitor is equal to the positive peak of V(4) (less
0.7 diode drop). The AC riding on the negative end, right end, is shifted down. The positive peak
of the waveform is clamped to 0 V (0.7 V) because the diode conducts on the positive peak.

1000 pF
+ - 1

+ 5 Vp-p
0 Voffset
0
- 1 kHz
(a)

-4.3 VDC

1000 pF
4
2
- +

4.3 VDC

+0
(b)

1000 pF
4
3
- +

+ 5V +

(c)

9.3 VDC

Figure 3.37: Clampers: (a) Positive peak clamped to 0 V. (b) Negative peak clamped to 0 V. (c)
Negative peak clamped to 5 V.
Suppose the polarity of the diode is reversed as in Figure 3.37 (b)? The diode conducts on the
negative peak of source V(4). The negative peak is clamped to 0 V (-0.7 V). See V(2) in Figure 3.38.

3.6.

CLAMPER CIRCUITS

117

*SPICE 03443.eps
V1 6 0 5
D1 6 3 diode
C1 4 3 1000p
D2 0 2 diode
C2 4 2 1000p
C3 4 1 1000p
D3 1 0 diode
V2 4 0 SIN(0 5 1k)
.model diode d
.tran 0.01m 5m
.end
Figure 3.38: V(4) source voltage 5 V peak used in all clampers. V(1) clamper output from Figure 3.37
(a). V(1,4) DC voltage on capacitor in Figure (a). V(2) clamper output from Figure (b). V(3)
clamper output from Figure (c).
The most general realization of the clamper is shown in Figure 3.37 (c) with the diode connected
to a DC reference. The capacitor still charges during the negative peak of the source. Note that the
polarities of the AC source and the DC reference are series aiding. Thus, the capacitor charges to
the sum to the two, 10 V DC (9.3 V). Coupling the 5 V peak sinewave across the capacitor yields
Figure 3.38 V(3), the sum of the charge on the capacitor and the sinewave. The negative peak
appears to be clamped to 5 V DC (4.3V), the value of the DC clamp reference (less diode drop).
Describe the waveform if the DC clamp reference is changed from 5 V to 10 V. The clamped
waveform will shift up. The negative peak will be clamped to 10 V (9.3). Suppose that the amplitude
of the sine wave source is increased from 5 V to 7 V? The negative peak clamp level will remain
unchanged. Though, the amplitude of the sinewave output will increase.
An application of the clamper circuit is as a DC restorer in composite video circuitry in
both television transmitters and receivers. An NTSC (US video standard) video signal white level
corresponds to minimum (12.5%) transmitted power. The video black level corresponds to a
high level (75% of transmitter power. There is a blacker than black level corresponding to 100%
transmitted power assigned to synchronization signals. The NTSC signal contains both video and
synchronization pulses. The problem with the composite video is that its average DC level varies
with the scene, dark vs light. The video itself is supposed to vary. However, the sync must always
peak at 100%. To prevent the sync signals from drifting with changing scenes, a DC restorer
clamps the top of the sync pulses to a voltage corresponding to 100% transmitter modulation. [2]
REVIEW:
A capacitively coupled signal alternates about its average DC level (0 V).

118

CHAPTER 3. DIODES AND RECTIFIERS


The signal out of a clamper appears the have one peak clamped to a DC voltage. Example:
The negative peak is clamped to 0 VDC, the waveform appears to be shifted upward. The
polarity of the diode determines which peak is clamped.
An application of a clamper, or DC restorer, is in clamping the sync pulses of composite video
to a voltage corresponding to 100% of transmitter power.

3.7

Voltage multipliers

A voltage multiplier is a specialized rectifier circuit producing an output which is theoretically an


integer times the AC peak input, for example, 2, 3, or 4 times the AC peak input. Thus, it is
possible to get 200 VDC from a 100 Vpeak AC source using a doubler, 400 VDC from a quadrupler.
Any load in a practical circuit will lower these voltages.
A voltage doubler application is a DC power supply capable of using either a 240 VAC or 120
VAC source. The supply uses a switch selected full-wave bridge to produce about 300 VDC from a
240 VAC source. The 120 V position of the switch rewires the bridge as a doubler producing about
300 VDC from the 120 VAC. In both cases, 300 VDC is produced. This is the input to a switching
regulator producing lower voltages for powering, say, a personal computer.
The half-wave voltage doubler in Figure 3.39 (a) is composed of two circuits: a clamper at (b)
and peak detector (half-wave rectifier) in Figure 3.28, which is shown in modified form in Figure 3.39
(c). C2 has been added to a peak detector (half-wave rectifier).

1000 pF

D1

C2

D2
5 Vp-p
1000 pF
0 Voffset
1 kHz 0

(a)

5V

2
C1

C2
D2

+
(b)

5V

+
+

C2
5V

D1

C1

+
10V

(c)

Figure 3.39: Half-wave voltage doubler (a) is composed of (b) a clamper and (c) a half-wave rectifier.
Referring to Figure 3.39 (b), C2 charges to 5 V (4.3 V considering the diode drop) on the negative
half cycle of AC input. The right end is grounded by the conducting D2. The left end is charged at
the negative peak of the AC input. This is the operation of the clamper.
During the positive half cycle, the half-wave rectifier comes into play at Figure 3.39 (c). Diode
D2 is out of the circuit since it is reverse biased. C2 is now in series with the voltage source. Note the
polarities of the generator and C2, series aiding. Thus, rectifier D1 sees a total of 10 V at the peak
of the sinewave, 5 V from generator and 5 V from C2. D1 conducts waveform v(1) (Figure 3.40),
charging C1 to the peak of the sine wave riding on 5 V DC (Figure 3.40 v(2)). Waveform v(2) is
the output of the doubler, which stabilizes at 10 V (8.6 V with diode drops) after a few cycles of
sinewave input.
The full-wave voltage doubler is composed of a pair of series stacked half-wave rectifiers. (Figure 3.41) The corresponding netlist is in Figure 3.41. The bottom rectifier charges C1 on the negative
half cycle of input. The top rectifier charges C2 on the positive halfcycle. Each capacitor takes on

3.7.

VOLTAGE MULTIPLIERS

119

*SPICE 03255.eps
C1 2 0 1000p
D1 1 2 diode
C2 4 1 1000p
D2 0 1 diode
V1 4 0 SIN(0 5 1k)
.model diode d
.tran 0.01m 5m
.end
Figure 3.40: Voltage doubler: v(4) input. v(1) clamper stage. v(2) half-wave rectifier stage, which
is the doubler output.
a charge of 5 V (4.3 V considering diode drop). The output at node 5 is the series total of C1 + C2
or 10 V (8.6 V with diode drops).

2
5 Vp-p
0 Voffset
1 kHz

D2

5
1000 pF

C2

3
1000 pF

C1

0
D1

*SPICE 03273.eps
*R1 3 0 100k
*R2 5 3 100k
D1 0 2 diode
D2 2 5 diode
C1 3 0 1000p
C2 5 3 1000p
V1 2 3 SIN(0 5 1k)
.model diode d
.tran 0.01m 5m
.end

Figure 3.41: Full-wave voltage doubler consists of two half-wave rectifiers operating on alternating
polarities.
Note that the output v(5) Figure 3.42 reaches full value within one cycle of the input v(2)
excursion.
Figure 3.43 illustrates the derivation of the full-wave doubler from a pair of opposite polarity halfwave rectifiers (a). The negative rectifier of the pair is redrawn for clarity (b). Both are combined
at (c) sharing the same ground. At (d) the negative rectifier is re-wired to share one voltage source
with the positive rectifier. This yields a 5 V (4.3 V with diode drop) power supply; though, 10

120

CHAPTER 3. DIODES AND RECTIFIERS

Figure 3.42: Full-wave voltage doubler: v(2) input, v(3)voltage at mid point, v(5) voltage at output
V is measurable between the two outputs. The ground reference point is moved so that +10 V is
available with respect to ground.
+5V

+5V

+5V

+10V

+5V

+
+5V

+
-

-5V
-5V

-5V

-5V

(a)

(b)

(c)

(d)

(e)

Figure 3.43: Full-wave doubler: (a) Pair of doublers, (b) redrawn, (c) sharing the ground, (d) share
the same voltage source. (e) move the ground point.
A voltage tripler (Figure 3.44) is built from a combination of a doubler and a half wave rectifier
(C3, D3). The half-wave rectifier produces 5 V (4.3 V) at node 3. The doubler provides another 10
V (8.4 V) between nodes 2 and 3. for a total of 15 V (12.9 V) at the output node 2 with respect to
ground. The netlist is in Figure 3.45.
Note that V(3) in Figure 3.45 rises to 5 V (4.3 V) on the first negative half cycle. Input v(4) is
shifted upward by 5 V (4.3 V) due to 5 V from the half-wave rectifier. And 5 V more at v(1) due
to the clamper (C2, D2). D1 charges C1 (waveform v(2)) to the peak value of v(1).

3.7.

VOLTAGE MULTIPLIERS

121

1000 pF

4
5 Vp-p
0 Voffset
1 kHz
(a)

C2

Doubler

2
D1

D2

C1
10V

1000 pF

10V
C3

Single stage rectifier

1000 pF
D3

5V

Figure 3.44: Voltage tripler composed of doubler stacked atop a single stage rectifier.

*SPICE 03283.eps
C3 3 0 1000p
D3 0 4 diode
C1 2 3 1000p
D1 1 2 diode
C2 4 1 1000p
D2 3 1 diode
V1 4 3 SIN(0 5 1k)
.model diode d
.tran 0.01m 5m
.end
Figure 3.45: Voltage tripler: v(3) half-wave rectifier, v(4) input+ 5 V, v(1) clamper, v(2) final
output.

122

CHAPTER 3. DIODES AND RECTIFIERS

A voltage quadrupler is a stacked combination of two doublers shown in Figure 3.46. Each
doubler provides 10 V (8.6 V) for a series total at node 2 with respect to ground of 20 V (17.2 V).
The netlist is in Figure 3.47.

1000 pF

4
5 Vp-p
0 Voffset
1 kHz

D1

C2
C1
D2

doubler 1
doubler 2
1000 pF
C22

1000 pF
3

10V
20V

C11
D22

1000 pF
D11

10V

Figure 3.46: Voltage quadrupler, composed of two doublers stacked in series, with output at node 2.
The waveforms of the quadrupler are shown in Figure 3.47. Two DC outputs are available: v(3),
the doubler output, and v(2) the quadrupler output. Some of the intermediate voltages at clampers
illustrate that the input sinewave (not shown), which swings by plusminus)5 V, is successively
clamped at higher levels: at v(5), v(4) and v(1). Strictly v(4) is not a clamper output. It is simply
the AC voltage source in series with the v(3) the doubler output. None the less, v(1) is a clamped
version of v(4)

*SPICE 03441.eps
*SPICE 03286.eps
C22 4 5 1000p
C11 3 0 1000p
D11 0 5 diode
D22 5 3 diode
C1 2 3 1000p
D1 1 2 diode
C2 4 1 1000p
D2 3 1 diode
V1 4 3 SIN(0 5 1k)
.model diode d
.tran 0.01m 5m
.end
Figure 3.47: Voltage quadrupler: DC voltage available at v(3) and v(2). Intermediate waveforms:
Clampers: v(5), v(4), v(1).
Some notes on voltage multipliers are in order at this point. The circuit parameters used in the

3.7.

VOLTAGE MULTIPLIERS

123

examples (V= 5 V 1 kHz, C=1000 pf) do not provide much current, microamps. Furthermore, load
resistors have been omitted. Loading reduces the voltages from those shown. If the circuits are to
be driven by a kHz source at low voltage, as in the examples, the capacitors are usually 0.1 to 1.0
F so that milliamps of current are available at the output. If the multipliers are driven from 50/60
Hz, the capacitor are a few hundred to a few thousand microfarads to provide hundreds of milliamps
of output current. If driven from line voltage, pay attention to the polarity and voltage ratings of
the capacitors.
Finally, any direct line driven power supply (no transformer) is dangerous to the experimenter
and line operated test equipment. Commercial direct driven supplies are safe because the hazardous
circuitry is in an enclosure to protect the user. When breadboarding these circuits with electrolytic
capacitors of any voltage, the capacitors will explode if the polarity is reversed. Such circuits should
be powered up behind a safety shield.
A voltage multiplier of cascaded half-wave doublers of arbitrary length is known as a CockcroftWalton multiplier as shown in Figure 3.48. This multiplier is used when a high voltage at low
current is required. The advantage over a conventional supply is that an expensive high voltage
transformer is not required at least not as high as the output.

1000 pF
99
1
5 Vp-p
0 Voffset
1 kHz
1000 pF

1000 pF

1000 pF

2
1000 pF

1000 pF

6
1000 pF

8
1000 pF

Figure 3.48: Cockcroft-Walton x8 voltage multiplier; output at v(8).


The pair of diodes and capacitors to the left of nodes 1 and 2 in Figure 3.48 constitute a half-wave
doubler. Rotating the diodes by 45o counterclockwise, and the bottom capacitor by 90o makes it
look like Figure 3.39 (a). Four of the doubler sections are cascaded to the right for a theoretical x8
multiplication factor. Node 1 has a clamper waveform (not shown), a sinewave shifted up by 1x (5
V). The other odd numbered nodes are sinewaves clamped to successively higher voltages. Node 2,
the output of the first doubler, is a 2x DC voltage v(2) in Figure 3.49. Successive even numbered
nodes charge to successively higher voltages: v(4), v(6), v(8)
Without diode drops, each doubler yields 2Vin or 10 V, considering two diode drops (10-1.4)=8.6
V is realistic. For a total of 4 doublers one expects 48.6=34.4 V out of 40 V. Consulting Figure 3.49,
v(2) is about right;however, v(8) is <30 V instead of the anticipated 34.4 V. The bane of the
Cockcroft-Walton multiplier is that each additional stage adds less than the previous stage. Thus,
a practical limit to the number of stages exist. It is possible to overcome this limitation with a
modification to the basic circuit. [3] Also note the time scale of 40 msec compared with 5 ms for
previous circuits. It required 40 msec for the voltages to rise to a terminal value for this circuit.
The netlist in Figure 3.49 has a .tran 0.010m 50m command to extend the simulation time to 50
msec; though, only 40 msec is plotted.
The Cockcroft-Walton multiplier serves as a more efficient high voltage source for photomultiplier
tubes requiring up to 2000 V. [3] Moreover, the tube has numerous dynodes, terminals requiring
connection to the lower voltage even numbered nodes. The series string of multiplier taps replaces

124

CHAPTER 3. DIODES AND RECTIFIERS


D1 7 8 diode
C1 8 6 1000p
D2 6 7 diode
C2 5 7 1000p
D3 5 6 diode
C3 4 6 1000p
D4 4 5 diode
C4 3 5 1000p
D5 3 4 diode
C5 2 4 1000p
D6 2 3 diode
D7 1 2 diode
C6 1 3 1000p
C7 2 0 1000p
C8 99 1 1000p
D8 0 1 diode
V1 99 0 SIN(0 5 1k)
.model diode d
.tran 0.01m 50m
.end
Figure 3.49: Cockcroft-Walton (x8) waveforms. Output is v(8).

a heat generating resistive voltage divider of previous designs.


An AC line operated Cockcroft-Walton multiplier provides high voltage to ion generators for
neutralizing electrostatic charge and for air purifiers.
REVIEW:
A voltage multiplier produces a DC multiple (2,3,4, etc) of the AC peak input voltage.
The most basic multiplier is a half-wave doubler.
The full-wave double is a superior circuit as a doubler.
A tripler is a half-wave doubler and a conventional rectifier stage (peak detector).
A quadrupler is a pair of half-wave doublers
A long string of half-wave doublers is known as a Cockcroft-Walton multiplier.

3.8

Inductor commutating circuits

A popular use of diodes is for the mitigation of inductive kickback: the pulses of high voltage
produced when direct current through an inductor is interrupted. Take, for example, this simple
circuit in Figure 3.50 with no protection against inductive kickback.

3.8.

INDUCTOR COMMUTATING CIRCUITS

125
V

(a)

(b)

(c)

off
(d)

on
off

Figure 3.50: Inductive kickback: (a) Switch open. (b) Switch closed, electron current flows from
battery through coil which has polarity matching battery. Magnetic field stores energy. (c) Switch
open, Current still flows in coil due to collapsing magnetic field. Note polarity change on coil
changed. (d) Coil voltage vs time.
When the pushbutton switch is actuated, current goes through the inductor, producing a magnetic field around it. When the switch is de-actuated, its contacts open, interrupting current through
the inductor, and causing the magnetic field to rapidly collapse. Because the voltage induced in a
coil of wire is directly proportional to the rate of change over time of magnetic flux (Faradays Law:
e = Nd/dt), this rapid collapse of magnetism around the coil produces a high voltage spike.
If the inductor in question is an electromagnet coil, such as in a solenoid or relay (constructed
for the purpose of creating a physical force via its magnetic field when energized), the effect of
inductive kickback serves no useful purpose at all. In fact, it is quite detrimental to the switch,
as it causes excessive arcing at the contacts, greatly reducing their service life. Of the practical
methods for mitigating the high voltage transient created when the switch is opened, none so simple
as the so-called commutating diode in Figure 3.51.

(a)

(b)

+
-

(c)

Figure 3.51: Inductive kickback with protection: (a) Switch open. (b)Switch closed, storing energy
in magnetic field. (c) Switch open, inductive kickback is shorted by diode.
In this circuit, the diode is placed in parallel with the coil, such that it will be reverse-biased
when DC voltage is applied to the coil through the switch. Thus, when the coil is energized, the
diode conducts no current in Figure 3.51 (b).
However, when the switch is opened, the coils inductance responds to the decrease in current by
inducing a voltage of reverse polarity, in an effort to maintain current at the same magnitude and
in the same direction. This sudden reversal of voltage polarity across the coil forward-biases the
diode, and the diode provides a current path for the inductors current, so that its stored energy is
dissipated slowly rather than suddenly in Figure 3.51 (c).
As a result, the voltage induced in the coil by its collapsing magnetic field is quite low: merely
the forward voltage drop of the diode, rather than hundreds of volts as before. Thus, the switch

126

CHAPTER 3. DIODES AND RECTIFIERS

contacts experience a voltage drop equal to the battery voltage plus about 0.7 volts (if the diode is
silicon) during this discharge time.
In electronics parlance, commutation refers to the reversal of voltage polarity or current direction.
Thus, the purpose of a commutating diode is to act whenever voltage reverses polarity, for example,
on an inductor coil when current through it is interrupted. A less formal term for a commutating
diode is snubber, because it snubs or squelches the inductive kickback.
A noteworthy disadvantage of this method is the extra time it imparts to the coils discharge.
Because the induced voltage is clamped to a very low value, its rate of magnetic flux change over time
is comparatively slow. Remember that Faradays Law describes the magnetic flux rate-of-change
(d/dt) as being proportional to the induced, instantaneous voltage (e or v ). If the instantaneous
voltage is limited to some low figure, then the rate of change of magnetic flux over time will likewise
be limited to a low (slow) figure.
If an electromagnet coil is snubbed with a commutating diode, the magnetic field will dissipate
at a relatively slow rate compared to the original scenario (no diode) where the field disappeared
almost instantly upon switch release. The amount of time in question will most likely be less than
one second, but it will be measurably slower than without a commutating diode in place. This may
be an intolerable consequence if the coil is used to actuate an electromechanical relay, because the
relay will possess a natural time delay upon coil de-energization, and an unwanted delay of even
a fraction of a second may wreak havoc in some circuits.
Unfortunately, one cannot eliminate the high-voltage transient of inductive kickback and maintain fast de-magnetization of the coil: Faradays Law will not be violated. However, if slow demagnetization is unacceptable, a compromise may be struck between transient voltage and time by
allowing the coils voltage to rise to some higher level (but not so high as without a commutating
diode in place). The schematic in Figure 3.52 shows how this can be done.

V
off

(d)
(e)

on

(c)
(a)

(b)

off

Figure 3.52: (a) Commutating diode with series resistor. (b) Voltage waveform. (c) Level with no
diode. (d) Level with diode, no resistor. (e) Compromise level with diode and resistor.

A resistor placed in series with the commutating diode allows the coils induced voltage to rise to
a level greater than the diodes forward voltage drop, thus hastening the process of de-magnetization.
This, of course, will place the switch contacts under greater stress, and so the resistor must be sized
to limit that transient voltage at an acceptable maximum level.

3.9.

3.9

DIODE SWITCHING CIRCUITS

127

Diode switching circuits

Diodes can perform switching and digital logic operations. Forward and reverse bias switch a diode
between the low and high impedance states, respectively. Thus, it serves as a switch.

3.9.1

Logic

Diodes can perform digital logic functions: AND, and OR. Diode logic was used in early digital
computers. It only finds limited application today. Sometimes it is convenient to fashion a single
logic gate from a few diodes.
V+

(a)

Y=1

V+

Y=0

(b)

A B Y
0 0 0
0 1 0
1 0 0
1 1 1

(c)

Figure 3.53: Diode AND gate


An AND gate is shown in Figure 3.53. Logic gates have inputs and an output (Y) which is a
function of the inputs. The inputs to the gate are high (logic 1), say 10 V, or low, 0 V (logic 0).
In the figure, the logic levels are generated by switches. If a switch is up, the input is effectively
high (1). If the switch is down, it connects a diode cathode to ground, which is low (0). The output
depends on the combination of inputs at A and B. The inputs and output are customarily recorded
in a truth table at (c) to describe the logic of a gate. At (a) all inputs are high (1). This is
recorded in the last line of the truth table at (c). The output, Y, is high (1) due to the V + on
the top of the resistor. It is unaffected by open switches. At (b) switch A pulls the cathode of the
connected diode low, pulling output Y low (0.7 V). This is recorded in the third line of the truth
table. The second line of the truth table describes the output with the switches reversed from (b).
Switch B pulls the diode and output low. The first line of the truth table recordes the Output=0
for both input low (0). The truth table describes a logical AND function. Summary: both inputs A
and B high yields a high (1) out.
A two input OR gate composed of a pair of diodes is shown in Figure ??. If both inputs are logic
low at (a) as simulated by both switches downward, the output Y is pulled low by the resistor.
This logic zero is recorded in the first line of the truth table at (c). If one of the inputs is high as at
(b), or the other input is high, or both inputs high, the diode(s) conduct(s), pulling the output Y
high. These results are reordered in the second through fourth lines of the truth table. Summary:
any input high is a high out at Y.
A backup battery may be OR-wired with a line operated DC power supply in Figure 3.54 (d)
to power a load, even during a power failure. With AC power present, the line supply powers the
load, assuming that it is a higher voltage than the battery. In the event of a power failure, the line
supply voltage drops to 0 V; the battery powers the load. The diodes must be is series with the
power sources to prevent a failed line supply from draining the battery, and to prevent it from over

128

CHAPTER 3. DIODES AND RECTIFIERS


V+

V+

Y=0

Y=1

line
operated
power
supply

A B Y
0 0 0
0 1 1
1 0 1
1 1 1

+
backup
battery

(a)

(b)

(c)

load

(d)
+

Figure 3.54: OR gate: (a) First line, truth table (TT). (b) Third line TT. (d) Logical OR of power
line supply and back-up battery.
charging the battery when line power is available. Does your PC computer retain its BIOS setting
when powered off? Does your VCR (video cassette recorder) retain the clock setting after a power
failure? (PC Yes, old VCR no, new VCR yes.)

3.9.2

Analog switch

Diodes can switch analog signals. A reverse biased diode appears to be an open circuit. A forward
biased diode is a low resistance conductor. The only problem is isolating the AC signal being switched
from the DC control signal. The circuit in Figure 3.55 is a parallel resonant network: resonant tuning
inductor paralleled by one (or more) of the switched resonator capacitors. This parallel LC resonant
circuit could be a preselector filter for a radio receiver. It could be the frequency determining network
of an oscillator (not shown). The digital control lines may be driven by a microprocessor interface.

decoupling
capacitor

RFC

RFC

RFC

switched
resonator
capacitor

RFC

switching
diode

resonant
tuning
inductor
large value
DC blocking
capacitor

+5V
digital control

Figure 3.55: Diode switch: A digital control signal (low) selects a resonator capacitor by forward
biasing the switching diode.
The large value DC blocking capacitor grounds the resonant tuning inductor for AC while blocking DC. It would have a low reactance compared to the parallel LC reactances. This prevents the

3.10.

ZENER DIODES

129

anode DC voltage from being shorted to ground by the resonant tuning inductor. A switched resonator capacitor is selected by pulling the corresponding digital control low. This forward biases
the switching diode. The DC current path is from +5 V through an RF choke (RFC), a switching
diode, and an RFC to ground via the digital control. The purpose of the RFC at the +5 V is to
keep AC out of the +5 V supply. The RFC in series with the digital control is to keep AC out of
the external control line. The decoupling capacitor shorts the little AC leaking through the RFC to
ground, bypassing the external digital control line.
With all three digital control lines high (+5 V), no switched resonator capacitors are selected
due to diode reverse bias. Pulling one or more lines low, selects one or more switched resonator capacitors, respectively. As more capacitors are switched in parallel with the resonant tuning inductor,
the resonant frequency decreases.
The reverse biased diode capacitance may be substantial compared with very high frequency or
ultra high frequency circuits. PIN diodes may be used as switches for lower capacitance.

3.10

Zener diodes

If we connect a diode and resistor in series with a DC voltage source so that the diode is forwardbiased, the voltage drop across the diode will remain fairly constant over a wide range of power
supply voltages as in Figure 3.56 (a).
According to the diode equation, the current through a forward-biased PN junction is proportional to e raised to the power of the forward voltage drop. Because this is an exponential function,
current rises quite rapidly for modest increases in voltage drop. Another way of considering this is
to say that voltage dropped across a forward-biased diode changes little for large variations in diode
current. In the circuit shown above, diode current is limited by the voltage of the power supply, the
series resistor, and the diodes voltage drop, which as we know doesnt vary much from 0.7 volts.
If the power supply voltage were to be increased, the resistors voltage drop would increase almost
the same amount, and the diodes voltage drop just a little. Conversely, a decrease in power supply
voltage would result in an almost equal decrease in resistor voltage drop, with just a little decrease
in diode voltage drop. In a word, we could summarize this behavior by saying that the diode is
regulating the voltage drop at approximately 0.7 volts.
Voltage regulation is a useful diode property to exploit. Suppose we were building some kind of
circuit which could not tolerate variations in power supply voltage, but needed to be powered by
a chemical battery, whose voltage changes over its lifetime. We could form a circuit as shown and
connect the circuit requiring steady voltage across the diode, where it would receive an unchanging
0.7 volts.
This would certainly work, but most practical circuits of any kind require a power supply voltage
in excess of 0.7 volts to properly function. One way we could increase our voltage regulation point
would be to connect multiple diodes in series, so that their individual forward voltage drops of
0.7 volts each would add to create a larger total. For instance, if we had ten diodes in series, the
regulated voltage would be ten times 0.7, or 7 volts in Figure 3.56 (b).
So long as the battery voltage never sagged below 7 volts, there would always be about 7 volts
dropped across the ten-diode stack.
If larger regulated voltages are required, we could either use more diodes in series (an inelegant
option, in my opinion), or try a fundamentally different approach. We know that diode forward
voltage is a fairly constant figure under a wide range of conditions, but so is reverse breakdown

130

CHAPTER 3. DIODES AND RECTIFIERS

0.7 V

(a)

7.0 V

(b)

Figure 3.56: Forward biased Si reference: (a) single diode, 0.7V, (b) 10-diodes in series 7.0V.
voltage, and breakdown voltage is typically much, much greater than forward voltage. If we reversed
the polarity of the diode in our single-diode regulator circuit and increased the power supply voltage
to the point where the diode broke down (could no longer withstand the reverse-bias voltage
impressed across it), the diode would similarly regulate the voltage at that breakdown point, not
allowing it to increase further as in Figure 3.57 (a).
50 V

Zener diode

+
150 V
(a)

Cathode
100 V

(b)

Anode

Figure 3.57: (a) Reverse biased Si small-signal diode breaks down at about 100V. (b) Symbol for
Zener diode.
Unfortunately, when normal rectifying diodes break down, they usually do so destructively.
However, it is possible to build a special type of diode that can handle breakdown without failing
completely. This type of diode is called a zener diode, and its symbol looks like Figure 3.57 (b).
When forward-biased, zener diodes behave much the same as standard rectifying diodes: they
have a forward voltage drop which follows the diode equation and is about 0.7 volts. In reversebias mode, they do not conduct until the applied voltage reaches or exceeds the so-called zener
voltage, at which point the diode is able to conduct substantial current, and in doing so will try to
limit the voltage dropped across it to that zener voltage point. So long as the power dissipated by
this reverse current does not exceed the diodes thermal limits, the diode will not be harmed.
Zener diodes are manufactured with zener voltages ranging anywhere from a few volts to hundreds
of volts. This zener voltage changes slightly with temperature, and like common carbon-composition
resistor values, may be anywhere from 5 percent to 10 percent in error from the manufacturers
specifications. However, this stability and accuracy is generally good enough for the zener diode to
be used as a voltage regulator device in common power supply circuit in Figure 3.58.
Please take note of the zener diodes orientation in the above circuit: the diode is reverse-biased,
and intentionally so. If we had oriented the diode in the normal way, so as to be forward-biased,
it would only drop 0.7 volts, just like a regular rectifying diode. If we want to exploit this diodes

3.10.

ZENER DIODES

131

+
12.6 V

Figure 3.58: Zener diode regulator circuit, Zener voltage = 12.6V).

reverse breakdown properties, we must operate it in its reverse-bias mode. So long as the power
supply voltage remains above the zener voltage (12.6 volts, in this example), the voltage dropped
across the zener diode will remain at approximately 12.6 volts.
Like any semiconductor device, the zener diode is sensitive to temperature. Excessive temperature will destroy a zener diode, and because it both drops voltage and conducts current, it produces
its own heat in accordance with Joules Law (P=IE). Therefore, one must be careful to design the
regulator circuit in such a way that the diodes power dissipation rating is not exceeded. Interestingly enough, when zener diodes fail due to excessive power dissipation, they usually fail shorted
rather than open. A diode failed in this manner is readily detected: it drops almost zero voltage
when biased either way, like a piece of wire.
Lets examine a zener diode regulating circuit mathematically, determining all voltages, currents,
and power dissipations. Taking the same form of circuit shown earlier, well perform calculations
assuming a zener voltage of 12.6 volts, a power supply voltage of 45 volts, and a series resistor value
of 1000 (well regard the zener voltage to be exactly 12.6 volts so as to avoid having to qualify all
figures as approximate in Figure 3.59 (a)
If the zener diodes voltage is 12.6 volts and the power supplys voltage is 45 volts, there will be
32.4 volts dropped across the resistor (45 volts - 12.6 volts = 32.4 volts). 32.4 volts dropped across
1000 gives 32.4 mA of current in the circuit. (Figure 3.59 (b))

32.4 V
1 k

1 k

+
45 V

+
12.6 V

(a)

32.4 mA

45 V

12.6 V

(b)

32.4 mA

Figure 3.59: (a) Zener Voltage regulator with 1000 resistor. (b) Calculation of voltage drips and
current.
Power is calculated by multiplying current by voltage (P=IE), so we can calculate power dissipations for both the resistor and the zener diode quite easily:

132

CHAPTER 3. DIODES AND RECTIFIERS

Presistor = (32.4 mA)(32.4 V)


Presistor = 1.0498 W
Pdiode = (32.4 mA)(12.6 V)
Pdiode = 408.24 mW
A zener diode with a power rating of 0.5 watt would be adequate, as would a resistor rated for
1.5 or 2 watts of dissipation.
If excessive power dissipation is detrimental, then why not design the circuit for the least amount
of dissipation possible? Why not just size the resistor for a very high value of resistance, thus severely
limiting current and keeping power dissipation figures very low? Take this circuit, for example, with
a 100 k resistor instead of a 1 k resistor. Note that both the power supply voltage and the diodes
zener voltage in Figure 3.60 are identical to the last example:
32.4 V
100 k

324 A
12.6 V

45 V

324 A

Figure 3.60: Zener regulator with 100 k resistor.


With only 1/100 of the current we had before (324 A instead of 32.4 mA), both power dissipation
figures should be 100 times smaller:

Presistor = (324 A)(32.4 V)


Presistor = 10.498 mW
Pdiode = (324 A)(12.6 V)
Pdiode = 4.0824 mW
Seems ideal, doesnt it? Less power dissipation means lower operating temperatures for both the
diode and the resistor, and also less wasted energy in the system, right? A higher resistance value
does reduce power dissipation levels in the circuit, but it unfortunately introduces another problem.
Remember that the purpose of a regulator circuit is to provide a stable voltage for another circuit.
In other words, were eventually going to power something with 12.6 volts, and this something will
have a current draw of its own. Consider our first regulator circuit, this time with a 500 load
connected in parallel with the zener diode in Figure 3.61.
If 12.6 volts is maintained across a 500 load, the load will draw 25.2 mA of current. In order
for the 1 k series dropping resistor to drop 32.4 volts (reducing the power supplys voltage of 45
volts down to 12.6 across the zener), it still must conduct 32.4 mA of current. This leaves 7.2 mA

3.10.

ZENER DIODES

133

32.4 V
1 k

32.4 mA

45 V

25.2 mA

12.6 V

7.2 mA
32.4 mA

Rload
500

25.2 mA

Figure 3.61: Zener regulator with 1000 series resistor and 500 load.
of current through the zener diode.
Now consider our power-saving regulator circuit with the 100 k dropping resistor, delivering
power to the same 500 load. What it is supposed to do is maintain 12.6 volts across the load, just
like the last circuit. However, as we will see, it cannot accomplish this task. (Figure 3.62)
44.776 V
447.76 A

100 k
447.76 A

0 A
224 mV

45 V
447.76 A

Rload
500

447.76 A

Figure 3.62: Zener non-regulator with 100 K series resistor with 500 load.
With the larger value of dropping resistor in place, there will only be about 224 mV of voltage
across the 500 load, far less than the expected value of 12.6 volts! Why is this? If we actually had
12.6 volts across the load, it would draw 25.2 mA of current, as before. This load current would have
to go through the series dropping resistor as it did before, but with a new (much larger!) dropping
resistor in place, the voltage dropped across that resistor with 25.2 mA of current going through it
would be 2,520 volts! Since we obviously dont have that much voltage supplied by the battery, this
cannot happen.
The situation is easier to comprehend if we temporarily remove the zener diode from the circuit
and analyze the behavior of the two resistors alone in Figure 3.63.
Both the 100 k dropping resistor and the 500 load resistance are in series with each other,
giving a total circuit resistance of 100.5 k. With a total voltage of 45 volts and a total resistance
of 100.5 k, Ohms Law (I=E/R) tells us that the current will be 447.76 A. Figuring voltage drops
across both resistors (E=IR), we arrive at 44.776 volts and 224 mV, respectively. If we were to
re-install the zener diode at this point, it would see 224 mV across it as well, being in parallel
with the load resistance. This is far below the zener breakdown voltage of the diode and so it will
not break down and conduct current. For that matter, at this low voltage the diode wouldnt
conduct even if it were forward-biased! Thus, the diode ceases to regulate voltage. At least 12.6

134

CHAPTER 3. DIODES AND RECTIFIERS

44.776 V
447.76 A

100 k
447.76 A

224 mV

45 V
447.76 A

Rload
500

447.76 A

Figure 3.63: Non-regulator with Zener removed.

volts must be dropped across to activate it.


The analytical technique of removing a zener diode from a circuit and seeing whether or not
enough voltage is present to make it conduct is a sound one. Just because a zener diode happens to
be connected in a circuit doesnt guarantee that the full zener voltage will always be dropped across
it! Remember that zener diodes work by limiting voltage to some maximum level; they cannot make
up for a lack of voltage.
In summary, any zener diode regulating circuit will function so long as the loads resistance is
equal to or greater than some minimum value. If the load resistance is too low, it will draw too much
current, dropping too much voltage across the series dropping resistor, leaving insufficient voltage
across the zener diode to make it conduct. When the zener diode stops conducting current, it can
no longer regulate voltage, and the load voltage will fall below the regulation point.
Our regulator circuit with the 100 k dropping resistor must be good for some value of load
resistance, though. To find this acceptable load resistance value, we can use a table to calculate
resistance in the two-resistor series circuit (no diode), inserting the known values of total voltage
and dropping resistor resistance, and calculating for an expected load voltage of 12.6 volts:

Rdropping
E

Rload

Total

12.6

45

I
R

Volts
Amps
Ohms

100 k

With 45 volts of total voltage and 12.6 volts across the load, we should have 32.4 volts across
Rdropping :

Rdropping

Rload

Total

32.4

12.6

45

I
R

Volts
Amps

100 k

Ohms

With 32.4 volts across the dropping resistor, and 100 k worth of resistance in it, the current
through it will be 324 A:

3.10.

ZENER DIODES

135

Rdropping

Rload

Total

32.4

12.6

45

324

Amps

100 k

Ohms

Volts

Ohms Law
I= E
R
Being a series circuit, the current is equal through all components at any given time:

Rdropping

Rload

Total

32.4

12.6

45

Volts

324

324

324

Amps

100 k

Ohms

Rule of series circuits:


ITotal = I1 = I2 = . . . In
Calculating load resistance is now a simple matter of Ohms Law (R = E/I), giving us 38.889
k:

Rdropping

Rload

Total

32.4

12.6

45

Volts

324

324

324

Amps

100 k

38.889 k

Ohms

Ohms Law
R= E
I
Thus, if the load resistance is exactly 38.889 k, there will be 12.6 volts across it, diode or no
diode. Any load resistance smaller than 38.889 k will result in a load voltage less than 12.6 volts,
diode or no diode. With the diode in place, the load voltage will be regulated to a maximum of 12.6
volts for any load resistance greater than 38.889 k.
With the original value of 1 k for the dropping resistor, our regulator circuit was able to
adequately regulate voltage even for a load resistance as low as 500 . What we see is a tradeoff
between power dissipation and acceptable load resistance. The higher-value dropping resistor gave
us less power dissipation, at the expense of raising the acceptable minimum load resistance value.
If we wish to regulate voltage for low-value load resistances, the circuit must be prepared to handle

136

CHAPTER 3. DIODES AND RECTIFIERS

higher power dissipation.


Zener diodes regulate voltage by acting as complementary loads, drawing more or less current as
necessary to ensure a constant voltage drop across the load. This is analogous to regulating the speed
of an automobile by braking rather than by varying the throttle position: not only is it wasteful,
but the brakes must be built to handle all the engines power when the driving conditions dont
demand it. Despite this fundamental inefficiency of design, zener diode regulator circuits are widely
employed due to their sheer simplicity. In high-power applications where the inefficiencies would
be unacceptable, other voltage-regulating techniques are applied. But even then, small zener-based
circuits are often used to provide a reference voltage to drive a more efficient amplifier circuit
controlling the main power.
Zener diodes are manufactured in standard voltage ratings listed in Table 3.1. The table Common zener diode voltages lists common voltages for 0.3W and 1.3W parts. The wattage corresponds
to die and package size, and is the power that the diode may dissipate without damage.
Table 3.1:
0.5W
2.7V 3.0V
5.1V 5.6V
10V
11V
20V
24V
1.3W
4.7V 5.1V
9.1V 10V
18V
20V
36V
39V
68V
75V

Common zener diode voltages


3.3V
6.2V
12V
27V

3.6V
6.8V
13V
30V

3.9V
7.5V
15V

4.3V
8.2V
16V

4.7V
9.1V
18V

5.6V
11V
22V
43V
100V

6.2V
12V
24V
47V
200V

6.8V
13V
27V
51V

7.5V
15V
30V
56V

8.2V
16V
33V
62V

Zener diode clipper: A clipping circuit which clips the peaks of waveform a approximately
the zener voltage of the diodes. The circuit of Figure 3.64 has two zeners connected series opposing
to symmetrically clip a waveform at nearly the Zener voltage. The resistor limits current drawn by
the zeners to a safe value.

1
1.0 k

20 Vp
0 Voffset
0 1 kHz

2
V(2)
output

*SPICE 03445.eps
D1 4 0 diode
D2 4 2 diode
R1 2 1 1.0k
V1 1 0 SIN(0 20 1k)
.model diode d bv=10
.tran 0.001m 2m
.end

Figure 3.64: Zener diode clipper:


The zener breakdown voltage for the diodes is set at 10 V by the diode model parameter bv=10

3.11.

SPECIAL-PURPOSE DIODES

137

in the spice net list in Figure 3.64. This causes the zeners to clip at about 10 V. The back-to-back
diodes clip both peaks. For a positive half-cycle, the top zener is reverse biased, breaking down at
the zener voltage of 10 V. The lower zener drops approximately 0.7 V since it is forward biased.
Thus, a more accurate clipping level is 10+0.7=10.7V. Similar negative half-cycle clipping occurs a
-10.7 V. (Figure 3.65) shows the clipping level at a little over 10 V.

Figure 3.65: Zener diode clipper: v(1) input is clipped at waveform v(2).
REVIEW:
Zener diodes are designed to be operated in reverse-bias mode, providing a relatively low,
stable breakdown, or zener voltage at which they being to conduct substantial reverse current.
A zener diode may function as a voltage regulator by acting as an accessory load, drawing
more current from the source if the voltage is too high, and less if it is too low.

3.11

Special-purpose diodes

3.11.1

Schottky diodes

Schottky diodes are constructed of a metal -to-N junction rather than a P-N semiconductor junction.
Also known as hot-carrier diodes, Schottky diodes are characterized by fast switching times (low
reverse-recovery time), low forward voltage drop (typically 0.25 to 0.4 volts for a metal-silicon
junction), and low junction capacitance.
The schematic symbol for a schottky diode is shown in Figure 3.66.
The forward voltage drop (VF ), reverse-recovery time (trr ), and junction capacitance (CJ ) of
Schottky diodes are closer to ideal than the average rectifying diode. This makes them well
suited for high-frequency applications. Unfortunately, though, Schottky diodes typically have lower

138

CHAPTER 3. DIODES AND RECTIFIERS


Anode
Cathode

Figure 3.66: Schottky diode schematic symbol.


forward current (IF ) and reverse voltage (VRRM and VDC ) ratings than rectifying diodes and are
thus unsuitable for applications involving substantial amounts of power. Though they are used in
low voltage switching regulator power supplies.
Schottky diode technology finds broad application in high-speed computer circuits, where the
fast switching time equates to high speed capability, and the low forward voltage drop equates to
less power dissipation when conducting.
Switching regulator power supplies operating at 100s of kHz cannot use conventional silicon
diodes as rectifiers because of their slow switching speed . When the signal applied to a diode
changes from forward to reverse bias, conduction continues for a short time, while carriers are being
swept out of the depletion region. Conduction only ceases after this tr reverse recovery time has
expired. Schottky diodes have a shorter reverse recovery time.
Regardless of switching speed, the 0.7 V forward voltage drop of silicon diodes causes poor
efficiency in low voltage supplies. This is not a problem in, say, a 10 V supply. In a 1 V supply the
0.7 V drop is a substantial portion of the output. One solution is to use a schottky power diode
which has a lower forward drop.

3.11.2

Tunnel diodes

Tunnel diodes exploit a strange quantum phenomenon called resonant tunneling to provide a negative
resistance forward-bias characteristics. When a small forward-bias voltage is applied across a tunnel
diode, it begins to conduct current. (Figure 3.67(b)) As the voltage is increased, the current increases
and reaches a peak value called the peak current (IP ). If the voltage is increased a little more, the
current actually begins to decrease until it reaches a low point called the valley current (I V ). If the
voltage is increased further yet, the current begins to increase again, this time without decreasing
into another valley. The schematic symbol for the tunnel diode shown in Figure 3.67(a).
Tunnel diode
Anode

Forward
current

IP

Cathode
IV

(a)

(b)

VP

VV

Forward voltage

(c)

Figure 3.67: Tunnel diode (a) Schematic symbol. (b) Current vs voltage plot (c) Oscillator.
The forward voltages necessary to drive a tunnel diode to its peak and valley currents are known
as peak voltage (VP ) and valley voltage (VV ), respectively. The region on the graph where current

3.11.

SPECIAL-PURPOSE DIODES

139

is decreasing while applied voltage is increasing (between VP and VV on the horizontal scale) is
known as the region of negative resistance.
Tunnel diodes, also known as Esaki diodes in honor of their Japanese inventor Leo Esaki, are
able to transition between peak and valley current levels very quickly, switching between high and
low states of conduction much faster than even Schottky diodes. Tunnel diode characteristics are
also relatively unaffected by changes in temperature.

Breakdown voltage (V)

1000

Ge

100

Si

Ga

As
Ga

10
ling

ne

tun

1
1014

1015
1016
Doping concentration (cm-3)

1017

1018

Figure 3.68: Reverse breakdown voltage versus doping level. After Sze [22]
Tunnel diodes are heavily doped in both the P and N regions, 1000 times the level in a rectifier.
This can be seen in Figure 3.68. Standard diodes are to the far left, zener diodes near to the left,
and tunnel diodes to the right of the dashed line. The heavy doping produces an unusually thin
depletion region. This produces an unusually low reverse breakdown voltage with high leakage. The
thin depletion region causes high capacitance. To overcome this, the tunnel diode junction area
must be tiny. The forward diode characteristic consists of two regions: a normal forward diode
characteristic with current rising exponentially beyond VF , 0.3 V for Ge, 0.7 V for Si. Between
0 V and VF is an additional negative resistance characteristic peak. This is due to quantum
mechanical tunneling involving the dual particle-wave nature of electrons. The depletion region is
thin enough compared with the equivalent wavelength of the electron that they can tunnel through.
They do not have to overcome the normal forward diode voltage VF . The energy level of the
conduction band of the N-type material overlaps the level of the valence band in the P-type region.
With increasing voltage, tunneling begins; the levels overlap; current increases, up to a point. As
current increases further, the energy levels overlap less; current decreases with increasing voltage.
This is the negative resistance portion of the curve.
Tunnel diodes are not good rectifiers, as they have relatively high leakage current when reversebiased. Consequently, they find application only in special circuits where their unique tunnel effect
has value. To exploit the tunnel effect, these diodes are maintained at a bias voltage somewhere
between the peak and valley voltage levels, always in a forward-biased polarity (anode positive, and

140

CHAPTER 3. DIODES AND RECTIFIERS

cathode negative).
Perhaps the most common application of a tunnel diode is in simple high-frequency oscillator
circuits as in Figure 3.67(c), where it allows a DC voltage source to contribute power to an LC
tank circuit, the diode conducting when the voltage across it reaches the peak (tunnel) level and
effectively insulating at all other voltages. The resistors bias the tunnel diode at a few tenths of a
volt centered on the negative resistance portion of the characteristic curve. The L-C resonant circuit
may be a section of waveguide for microwave operation. Oscillation to 5 GHz is possible.
At one time the tunnel diode was the only solid-state microwave amplifier available. Tunnel
diodes were popular starting in the 1960s. They were longer lived than traveling wave tube amplifiers, an important consideration in satellite transmitters. Tunnel diodes are also resistant to
radiation because of the heavy doping. Today various transistors operate at microwave frequencies.
Even small signal tunnel diodes are expensive and difficult to find today. There is one remaining
manufacturer of germanium tunnel diodes, and none for silicon devices. They are sometimes used
in military equipment because they are insensitive to radiation and large temperature changes.
There has been some research involving possible integration of silicon tunnel diodes into CMOS
integrated circuits. They are thought to be capable of switching at 100 GHz in digital circuits. The
sole manufacturer of germanium devices produces them one at a time. A batch process for silicon
tunnel diodes must be developed, then integrated with conventional CMOS processes. [21]
The Esaki tunnel diode should not be confused with the resonant tunneling diode (page 80), of
more complex construction from compound semiconductors. The RTD is a more recent development
capable of higher speed.

3.11.3

Light-emitting diodes

Diodes, like all semiconductor devices, are governed by the principles described in quantum physics.
One of these principles is the emission of specific-frequency radiant energy whenever electrons fall
from a higher energy level to a lower energy level. This is the same principle at work in a neon
lamp, the characteristic pink-orange glow of ionized neon due to the specific energy transitions of its
electrons in the midst of an electric current. The unique color of a neon lamps glow is due to the fact
that its neon gas inside the tube, and not due to the particular amount of current through the tube
or voltage between the two electrodes. Neon gas glows pinkish-orange over a wide range of ionizing
voltages and currents. Each chemical element has its own signature emission of radiant energy
when its electrons jump between different, quantized energy levels. Hydrogen gas, for example,
glows red when ionized; mercury vapor glows blue. This is what makes spectrographic identification
of elements possible.
Electrons flowing through a PN junction experience similar transitions in energy level, and emit
radiant energy as they do so. The frequency of this radiant energy is determined by the crystal
structure of the semiconductor material, and the elements comprising it. Some semiconductor
junctions, composed of special chemical combinations, emit radiant energy within the spectrum of
visible light as the electrons change energy levels. Simply put, these junctions glow when forward
biased. A diode intentionally designed to glow like a lamp is called a light-emitting diode, or LED.
Forward biased silicon diodes give off heat as electron and holes from the N-type and P-type
regions, respectively, recombine at the junction. In a forward biased LED, the recombination of
electrons and holes in the active region in Figure 3.69 (c) yields photons. This process is known as
electroluminescence. To give off photons, the potential barrier through which the electrons fall must

3.11.

SPECIAL-PURPOSE DIODES

141

be higher than for a silicon diode. The forward diode drop can range to a few volts for some color
LEDs.
Diodes made from a combination of the elements gallium, arsenic, and phosphorus (called
gallium-arsenide-phosphide) glow bright red, and are some of the most common LEDs manufactured. By altering the chemical constituency of the PN junction, different colors may be obtained.
Some of the currently available colors other than red are green, blue, and infra-red (invisible light
at a frequency lower than red). Other colors may be obtained by combining two or more primarycolor (red, green, and blue) LEDs together in the same package, sharing the same optical lens. For
instance, a yellow LED may be made by merging a red LED with a green LED.
The schematic symbol for an LED is a regular diode shape inside of a circle, with two small
arrows pointing away (indicating emitted light), shown in Figure 3.69.

+
Anode

long

Cathode
(a)

short
(b)

flat
(c)

p-type
active region
n-type
substrate
electron
hole

Figure 3.69: LED, Light Emitting Diode: (a) schematic symbol. (b) Flat side and short lead of
device correspond to cathode. (c) Cross section of Led die.
This notation of having two small arrows pointing away from the device is common to the
schematic symbols of all light-emitting semiconductor devices. Conversely, if a device is lightactivated (meaning that incoming light stimulates it), then the symbol will have two small arrows
pointing toward it. LEDs can sense light. They generate a small voltage when exposed to light, much
like a solar cell on a small scale. This property can be gainfully applied in a variety of light-sensing
circuits.
Because LEDs are made of different chemical substances than silicon diodes, their forward voltage
drops will be different. Typically, LEDs have much larger forward voltage drops than rectifying
diodes, anywhere from about 1.6 volts to over 3 volts, depending on the color. Typical operating
current for a standard-sized LED is around 20 mA. When operating an LED from a DC voltage source
greater than the LEDs forward voltage, a series-connected dropping resistor must be included to
prevent full source voltage from damaging the LED. Consider the example circuit in Figure 3.70 (a)
using a 6 V source.
With the LED dropping 1.6 volts, there will be 4.4 volts dropped across the resistor. Sizing the
resistor for an LED current of 20 mA is as simple as taking its voltage drop (4.4 volts) and dividing
by circuit current (20 mA), in accordance with Ohms Law (R=E/I). This gives us a figure of 220 .
Calculating power dissipation for this resistor, we take its voltage drop and multiply by its current
(P=IE), and end up with 88 mW, well within the rating of a 1/8 watt resistor. Higher battery
voltages will require larger-value dropping resistors, and possibly higher-power rating resistors as
well. Consider the example in Figure 3.70 (b) for a supply voltage of 24 volts:
Here, the dropping resistor must be increased to a size of 1.12 k to drop 22.4 volts at 20 mA

142

CHAPTER 3. DIODES AND RECTIFIERS


Rdropping

220

Rdropping
Red LED,

+
24 V

VF = 1.6 V typical

6V

IF = 20 mA typical
(a)

1.12 k

(b)

Figure 3.70: Setting LED current at 20 ma. (a) for a 6 V source, (b) for a 24 V source.
so that the LED still receives only 1.6 volts. This also makes for a higher resistor power dissipation:
448 mW, nearly one-half a watt of power! Obviously, a resistor rated for 1/8 watt power dissipation
or even 1/4 watt dissipation will overheat if used here.
Dropping resistor values need not be precise for LED circuits. Suppose we were to use a 1 k
resistor instead of a 1.12 k resistor in the circuit shown above. The result would be a slightly
greater circuit current and LED voltage drop, resulting in a brighter light from the LED and slightly
reduced service life. A dropping resistor with too much resistance (say, 1.5 k instead of 1.12 k)
will result in less circuit current, less LED voltage, and a dimmer light. LEDs are quite tolerant of
variation in applied power, so you need not strive for perfection in sizing the dropping resistor.
Multiple LEDs are sometimes required, say in lighting. If LEDs are operated in parallel, each
must have its own current limiting resistor as in Figure ?? (a) to ensure currents dividing more
equally. However, it is more efficient to operate LEDs in series (Figure 3.71 (b)) with a single
dropping resistor. As the number of series LEDs increases the series resistor value must decrease to
maintain current, to a point. The number of LEDs in series (Vf ) cannot exceed the capability of
the power supply. Multiple series strings may be employed as in Figure 3.71 (c).
In spite of equalizing the currents in multiple LEDs, the brightness of the devices may not match
due to variations in the individual parts. Parts can be selected for brightness matching for critical
applications.

220

6V

140

(a)

6V

140
6V

(b)

(c)

Figure 3.71: Multiple LEDs: (a) In parallel, (b) in series, (c) series-parallel
Also because of their unique chemical makeup, LEDs have much, much lower peak-inverse voltage
(PIV) ratings than ordinary rectifying diodes. A typical LED might only be rated at 5 volts in
reverse-bias mode. Therefore, when using alternating current to power an LED, connect a protective
rectifying diode anti-parallel with the LED to prevent reverse breakdown every other half-cycle as

3.11.

SPECIAL-PURPOSE DIODES

143

in Figure 3.72 (a).


Rdropping

Red LED,
1.12 k
24 V

(a)

rectifying diode

VF = 1.6 V typical
IF = 20 mA typical
VR = 5 V maximum

(b)

0.22
F
240 V

1.2
k

Figure 3.72: Safely driving an LED with AC: (a) from 24 VAC, (b) from 240 VAC.
If the LED is driven from a 240 VAC source, the Figure 3.72 (a) voltage source is increased from
24 VAC to 240 VAC, the resistor from 1.12 k to 12 k. The power dissipated in the 12 k resistor
is an unattractive 4.8 watts.
P = VI = (240 V)(20 mA) = 4.8 watt
A potential solution is to replace the 12 k resistor with a non-dissipative 12 k capacitive
reactance. This would be Figure 3.72 (b) with the resistor shorted. That circuit at (b), missing the
resistor, was published in an electrical engineering journal. This author constructed the circuit. It
worked the first time it was powered on, but not thereafter upon power on. Each time it was
powered on, it got dimmer until it failed completely. Why? If power on occurs near a zero
crossing of the AC sinewave, the circuit works. However, if powered on at, say, the peak of the
sinewave, the voltage rises abruptly from zero to the peak. Since the current through the capacitor
is i = C(dv/dt), the current spikes to a very large value exceeding the surge current rating of the
LED, destroying it.
The solution is to design a capacitor for the continuous current of the LED, and a series resistor
to limit current during power on to the surge current rating of the LED. Often the surge current
rating of an LED is ten times higher than the continuous current rating. (Though, this is not true
of high current illumination grade LEDs.) We calculate a capacitor to supply 20 mA continuous
current, then select a resistor having resistance of 1/10 th the capacitive reactance.
I = 20 mA
Xc = (240 V) / (20 mA) = 12 k
Xc = 1/2fc
C = 1/2Xc = 1/260(12 k = 0.22 F
R = (0.10)Xc = (0.10)(12k) = 1.2 k
P = I2 R = (20 mA)2 (1.2 k) = 0.48 watt
The resistor limits the LED current to 200 mA during the power on surge. Thereafter it passes
20 mA as limited by the capacitor. The 1.2 kOMEGA resistor dissipates 0.48 watts compared with
4.8 watts for the 12 k resistor circuit.
What component values would be required to operate the circuit on 120 VAC? One solution
is to use the 240 VAC circuit on 120 VAC with no change in component values, halving the LED
continuous current to 10 mA. If operation at 20 mA is required, double the capacitor value and
halve the resistor value.

144

CHAPTER 3. DIODES AND RECTIFIERS

The anti-parallel diodes in Figure 3.72 can be replaced with an anti-parallel LED. The resulting
pair of anti-parallel LEDs illuminate on alternating half-cycles of the AC sinewave. This configuration draws 20 ma, splitting it equally between the LEDs on alternating AC half cycles. Each LED
only receives 10 ma due to this sharing. The same is true of the LED anti-parallel combination with
a rectifier. The LED only receives 10 ma. If 20 ma was required for the LED(s), The capacitor value
in F could be doubled and the resistor halved.
The forward voltage drop of LEDs is inversely proportional to the wavelength (). As wavelength
decreases going from infrared to visible colors to ultraviolet, Vf increases. While this trend is most
obvious in the various devices from a single manufacturer, The voltage range for a particular color
LED from various manufacturers varies. This range of voltages is shown in Table 3.2.
Table 3.2: Optical and electrical properties of LEDs
LED
nm (= 10 9 m) Vf (from) Vf (to)
infrared
940
1.2
1.7
red
660
1.5
2.4
orange
602-620
2.1
2.2
yellow, green
560-595
1.7
2.8
white, blue, violet 3
4
ultraviolet
370
4.2
4.8
As lamps, LEDs are superior to incandescent bulbs in many ways. First and foremost is efficiency:
LEDs output far more light power per watt of electrical input than an incandescent lamp. This is
a significant advantage if the circuit in question is battery-powered, efficiency translating to longer
battery life. Second is the fact that LEDs are far more reliable, having a much greater service life
than an incandescent lamps. This advantage is because LEDs are cold devices: they operate at
much cooler temperatures than an incandescent lamp with a white-hot metal filament, susceptible
to breakage from mechanical and thermal shock. Third is the high speed at which LEDs may be
turned on and off. This advantage is also due to the cold operation of LEDs: they dont have to
overcome thermal inertia in transitioning from off to on or vice versa. For this reason, LEDs are
used to transmit digital (on/off) information as pulses of light, conducted in empty space or through
fiber-optic cable, at very high rates of speed (millions of pulses per second).
LEDs excel in monochromatic lighting applications like traffic signals and automotive tail lights.
Incandescents are abysmal in this application since they require filtering, decreasing efficiency. LEDs
do not need filtering.
One major disadvantage of using LEDs as sources of illumination is their monochromatic (singlecolor) emission. No one wants to read a book under the light of a red, green, or blue LED. However,
if used in combination, LED colors may be mixed for a more broad-spectrum glow. A new broad
spectrum light source is the white LED. While small white panel indicators have been available for
many years, illumination grade devices are still in development.
A white LED is a blue LED exciting a phosphor which emits yellow light. The blue plus yellow
approximates white light. The nature of the phosphor determines the characteristics of the light. A
red phosphor may be added to improve the quality of the yellow plus blue mixture at the expense
of efficiency. Table 3.3 compares white illumination LEDs to expected future devices and other
conventional lamps. Efficiency is measured in lumens of light output per watt of input power. If
the 50 lumens/watt device can be improved to 100 lumens/watt, white LEDs will be comparable to

3.11.

SPECIAL-PURPOSE DIODES

Lamp type
White LED
White LED, future
Incandescent
Halogen
Compact fluorescent
Sodium vapor, lp
Mercury vapor

Table 3.3: Efficiency of lighting


Efficiency lumen/watt Life hrs
35
100,000
100
100,000
12
1000
15-17
2000
50-100
10,000
70-200
20,000
13-48
18,000

145

notes
costly
R&D target
inexpensive
high quality light
cost effective
outdoor
outdoor

compact fluorescent lamps in efficiency.

3.11.4

Laser diodes

The laser diode is a further development upon the regular light-emitting diode, or LED. The term
laser itself is actually an acronym, despite the fact its often written in lower-case letters. Laser
stands for Light Amplification by Stimulated Emission of Radiation, and refers to another strange
quantum process whereby characteristic light emitted by electrons falling from high-level to low-level
energy states in a material stimulate other electrons in a substance to make similar jumps, the
result being a synchronized output of light from the material. This synchronization extends to the
actual phase of the emitted light, so that all light waves emitted from a lasing material are not
just the same frequency (color), but also the same phase as each other, so that they reinforce one
another and are able to travel in a very tightly-confined, nondispersing beam. This is why laser light
stays so remarkably focused over long distances: each and every light wave coming from the laser is
in step with each other.

"white"
light
source
(a)
monochromatic
light
source
(b)
laser
light
source
(c)

Figure 3.73: (a) White light of many wavelengths. (b) Mono-chromatic LED light, a single wavelength. (c) Phase coherent laser light.

146

CHAPTER 3. DIODES AND RECTIFIERS

Incandescent lamps produce white (mixed-frequency, or mixed-color) light as in Figure 3.73


(a). Regular LEDs produce monochromatic light: same frequency (color), but different phases,
resulting in similar beam dispersion in Figure 3.73 (b). Laser LEDs produce coherent light: light
that is both monochromatic (single-color) and monophasic (single-phase), resulting in precise beam
confinement as in Figure 3.73 (c).
Laser light finds wide application in the modern world: everything from surveying, where a
straight and nondispersing light beam is very useful for precise sighting of measurement markers, to
the reading and writing of optical disks, where only the narrowness of a focused laser beam is able
to resolve the microscopic pits in the disks surface comprising the binary 1s and 0s of digital
information.
Some laser diodes require special high-power pulsing circuits to deliver large quantities of
voltage and current in short bursts. Other laser diodes may be operated continuously at lower
power. In the continuous laser, laser action occurs only within a certain range of diode current,
necessitating some form of current-regulator circuit. As laser diodes age, their power requirements
may change (more current required for less output power), but it should be remembered that lowpower laser diodes, like LEDs, are fairly long-lived devices, with typical service lives in the tens of
thousands of hours.

3.11.5

Photodiodes

A photodiode is a diode optimized to produce an electron current flow in response to irradiation by


ultraviolet, visible, or infrared light. Silicon is most often used to fabricate photodiodes; though,
germanium and gallium arsenide can be used. The junction through which light enters the semiconductor must be thin enough to pass most of the light on to the active region (depletion region)
where light is converted to electron hole pairs.
In Figure 3.74 a shallow P-type diffusion into an N-type wafer produces a PN junction near the
surface of the wafer. The P-type layer needs to be thin to pass as much light as possible. A heavy
N+ diffusion on the back of the wafer makes contact with metalization. The top metalization may
be a fine grid of metallic fingers on the top of the wafer for large cells. In small photodiodes, the
top contact might be a sole bond wire contacting the bare P-type silicon top.

+
+

top metal contact


p diffusion
depletion region
n type

n+ contact region
bottom metal contact

Figure 3.74: Photodiode: Schematic symbol and cross section.

3.11.

SPECIAL-PURPOSE DIODES

147

Light entering the top of the photodiode stack fall off exponentially in with depth of the silicon.
A thin top P-type layer allows most photons to pass into the depletion region where electron-hole
pairs are formed. The electric field across the depletion region due to the built in diode potential
causes electrons to be swept into the N-layer, holes into the P-layer. Actually electron-hole pairs
may be formed in any of the semiconductor regions. However, those formed in the depletion region
are most likely to be separated into the respective N and P-regions. Many of the electron-hole
pairs formed in the P and N-regions recombine. Only a few do so in the depletion region. Thus,
a few electron-hole pairs in the N and P-regions, and most in the depletion region contribute to
photocurrent, that current resulting from light falling on the photodiode.
The voltage out of a photodiode may be observed. However, operation in this photovoltaic (PV)
mode is not linear over a large dynamic range. Though, it is sensitive and low noise at low frequencies,
< 100 kHz. The preferred mode of operation is often photocurrent (PC) mode because the current
is linearly proportional to light flux over several decades of intensity, and higher frequency response
can be achieved. PC mode is achieved with reverse bias or zero bias on the photodiode. A current
amplifier (transimpedance amplifier) should be used with a photodiode in PC mode. Linearity and
PC mode are achieved as long as the diode does not become forward biased.
High speed operation is often required of photodiodes, as opposed to solar cells. Speed is a
function of diode capacitance, which can be minimized by decreasing cell area. Thus, a sensor for a
high speed fiber optic link will use an area no larger than necessary, say 1 mm 2 . Capacitance may
also be decreased by increasing the thickness of the depletion region, in the manufacturing process
or by increasing the reverse bias on the diode.
PIN diode The p-i-n diode or PIN diode is a photodiode with an intrinsic layer between the
P and N-regions as in Figure 3.75. The P-Intrinsic-N structure increases the distance between
the P and N conductive layers, decreasing capacitance, increasing speed. The volume of the photo
sensitive region also increases, enhancing conversion efficiency. The bandwidth can extend to 10s
of gHz. PIN photodiodes are the preferred for high sensitivity, and high speed at moderate cost.

+
top metal contact
p diffusion
intrinsic region
(larger depletion
region)
n type
n+ contact region
bottom metal contact

Figure 3.75: PIN photodiode: The intrinsic region increases the thickness of the depletion region.
Avalanche photo diode:An avalanche photodiode (APD)designed to operate at high reverse

148

CHAPTER 3. DIODES AND RECTIFIERS

bias exhibits an electron multiplier effect analogous to a photomultiplier tube. The reverse bias can
run from 10s of volts to nearly 2000 V. The high level of reverse bias accelerates photon created
electron-hole pairs in the intrinsic region to a high enough velocity to free additional carriers from
collisions with the crystal lattice. Thus, many electrons per photon result. The motivation for the
APD is to achieve amplification within the photodiode to overcome noise in external amplifiers.
This works to some extent. However, the APD creates noise of its own. At high speed the APD
is superior to a PIN diode amplifier combination, though not for low speed applications. APDs
are expensive, roughly the price of a photomultiplier tube. So, they are only competitive with PIN
photodiodes for niche applications. One such application is single photon counting as applied to
nuclear physics.

3.11.6

Solar cells

A photodiode optimized for efficiently delivering power to a load is the solar cell. It operates
in photovoltaic mode (PV) because it is forward biased by the voltage developed across the load
resistance.
Monocrystalline solar cells are manufactured in a process similar to semiconductor processing.
This involves growing a single crystal boule from molten high purity silicon (P-type), though, not as
high purity as for semiconductors. The boule is diamond sawed or wire sawed into wafers. The ends
of the boule must be discarded or recycled, and silicon is lost in the saw kerf. Since modern cells
are nearly square, silicon is lost in squaring the boule. Cells may be etched to texture (roughen) the
surface to help trap light within the cell. Considerable silicon is lost in producing the 10 or 15 cm
square wafers. These days (2007) it is common for solar cell manufacturer to purchase the wafers at
this stage from a supplier to the semiconductor industry.
P-type Wafers are loaded back-to-back into fused silica boats exposing only the outer surface to
the N-type dopant in the diffusion furnace. The diffusion process forms a thin n-type layer on the
top of the cell. The diffusion also shorts the edges of the cell front to back. The periphery must
be removed by plasma etching to unshort the cell. Silver and or aluminum paste is screened on the
back of the cell, and a silver grid on the front. These are sintered in a furnace for good electrical
contact. (Figure 3.76)
The cells are wired in series with metal ribbons. For charging 12 V batteries, 36 cells at approximately 0.5 V are vacuum laminated between glass, and a polymer metal back. The glass may have
a textured surface to help trap light.
The ultimate commercial high efficiency (21.5%) single crystal silicon solar cells have all contacts
on the back of the cell. The active area of the cell is increased by moving the top (-) contact
conductors to the back of the cell. The top (-) contacts are normally made to the N-type silicon on
top of the cell. In Figure 3.77 the (-) contacts are made to N+ diffusions on the bottom interleaved
with (+) contacts. The top surface is textured to aid in trapping light within the cell.. [18]
Multicyrstalline silicon cells start out as molten silicon cast into a rectangular mold. As the
silicon cools, it crystallizes into a few large (mm to cm sized) randomly oriented crystals instead of
a single one. The remainder of the process is the same as for single crystal cells. The finished cells
show lines dividing the individual crystals, as if the cells were cracked. The high efficiency is not
quite as high as single crystal cells due to losses at crystal grain boundaries. The cell surface cannot
be roughened by etching due to the random orientation of the crystals. However, an antireflectrive
coating improves efficiency. These cells are competitive for all but space applications.

3.11.

SPECIAL-PURPOSE DIODES

149

top metal contact


N diffusion
depletion region
P type wafer
bottom metal
contact

+
Figure 3.76: Silicon Solar cell

Antireflectrive coating
Silicon dioxide passivation
N-type diffusion
P-type wafer
N+ diffusion
- contact
P+ diffusion
+ contact
N+ diffusion
- contact

Figure 3.77: High efficiency solar cell with all contacts on the back. Adapted from Figure 1 [18]

150

CHAPTER 3. DIODES AND RECTIFIERS

Three layer cell: The highest efficiency solar cell is a stack of three cells tuned to absorb
different portions of the solar spectrum. Though three cells can be stacked atop one another, a
monolithic single crystal structure of 20 semiconductor layers is more compact. At 32 % efficiency,
it is now (2007) favored over silicon for space application. The high cost prevents it from finding
many earth bound applications other than concentrators based on lenses or mirrors.
Intensive research has recently produced a version enhanced for terrestrial concentrators at 400
- 1000 suns and 40.7% efficiency. This requires either a big inexpensive Fresnel lens or reflector and
a small area of the expensive semiconductor. This combination is thought to be competitive with
inexpensive silicon cells for solar power plants. [8] [23]
Metal organic chemical vapor deposition (MOCVD) deposits the layers atop a P-type germanium
substrate. The top layers of N and P-type gallium indium phosphide (GaInP) having a band gap
of 1.85 eV, absorbs ultraviolet and visible light. These wavelengths have enough energy to exceed
the band gap. Longer wavelengths (lower energy) do not have enough energy to create electron-hole
pairs, and pass on through to the next layer. A gallium arsenide layers having a band gap of 1.42
eV, absorbs near infrared light. Finally the germanium layer and substrate absorb far infrared. The
series of three cells produce a voltage which is the sum of the voltages of the three cells. The voltage
developed by each material is 0.4 V less than the band gap energy listed in Table 3.4. For example,
for GaInP: 1.8 eV/e - 0.4 V = 1.4 V. For all three the voltage is 1.4 V + 1.0 V + 0.3 V = 2.7 V. [4]
Table 3.4: High efficiency triple layer solar cell.
Layer
Band gap Light absorbed
Gallium indium phosphide 1.8 eV
UV, visible
Gallium arsenide
1.4 eV
near infrared
Germanium
0.7 eV
far infrared
Crystalline solar cell arrays have a long useable life. Many arrays are guaranteed for 25 years, and
believed to be good for 40 years. They do not suffer initial degradation compared with amorphous
silicon.
Both single and multicrystalline solar cells are based on silicon wafers. The silicon is both the
substrate and the active device layers. Much silicon is consumed. This kind of cell has been around
for decades, and takes approximately 86% of the solar electric market. For further information about
crystalline solar cells see Honsberg. [7]
Amorphous silicon thin film solar cells use tiny amounts of the active raw material, silicon.
Approximately half the cost of conventional crystalline solar cells is the solar cell grade silicon. The
thin film deposition process reduces this cost. The downside is that efficiency is about half that of
conventional crystalline cells. Moreover, efficiency degrades by 15-35% upon exposure to sunlight.
A 7% efficient cell soon ages to 5% efficiency. Thin film amorphous silicon cells work better than
crystalline cells in dim light. They are put to good use in solar powered calculators.
Non-silicon based solar cells make up about 7% of the market. These are thin-film polycrystalline
products. Various compound semiconductors are the subject of research and development. Some
non-silicon products are in production. Generally, the efficiency is better than amorphous silicon,
but not nearly as good as crystalline silicon.
Cadmium telluride as a polycrystalline thin film on metal or glass can have a higher efficiency
than amorphous silicon thin films. If deposited on metal, that layer is the negative contact to the
cadmium telluride thin film. The transparent P-type cadmium sulfide atop the cadmium telluride

3.11.

SPECIAL-PURPOSE DIODES

151

serves as a buffer layer. The positive top contact is transparent, electrically conductive fluorine
doped tin oxide. These layers may be laid down on a sacrificial foil in place of the glass in the
process in the following pargraph. The sacrificial foil is removed after the cell is mounted to a
permanent substrate.

glass substrate
n Tin oxide
cadmium suflide
p cadmium telluride
(phosphorus doped)
p+ lead telluride
metal substrate
metal contact

Figure 3.78: Cadmium telluride solar cell on glass or metal.


A process for depositing cadmium telluride on glass begins with the deposition of N-type transparent, electrically conducive, tin oxide on a glass substrate. The next layer is P-type cadmium
telluride; though, N-type or intrinsic may be used. These two layers constitute the NP junction. A
P+ (heavy P-type) layer of lead telluride aids in establishing a low resistance contact. A metal layer
makes the final contact to the lead telluride. These layers may be laid down by vacuum deposition, chemical vapor deposition (CVD), screen printing, electrodeposition, or atmospheric pressure
chemical vapor deposition (APCVD) in helium. [10]
A variation of cadmium telluride is mercury cadmium telluride. Having lower bulk resistance
and lower contact resistance improves efficiency over cadmium telluride.
Cadmium Indium Gallium diSelenide: A most promising thin film solar cell at this time
(2007) is manufactured on a ten inch wide roll of flexible polyimide Cadmium Indium Gallium
diSelenide (CIGS). It has a spectacular efficiency of 10%. Though, commercial grade crystalline
silicon cells surpassed this decades ago, CIGS should be cost competitive. The deposition processes
are at a low enough temperature to use a polyimide polymer as a substrate instead of metal or glass.
The CIGS is manufactured in a roll to roll process, which should drive down costs.
REVIEW:
Most solar cells are silicon single crystal or multicrystal because of their good efficiency and
moderate cost.
Less efficient thin films of various amorphous or polycrystalline materials comprise the rest of
the market.
Table ?? compares selected solar cells.

152

CHAPTER 3. DIODES AND RECTIFIERS

top contact
N-type transparent
conductor
buffer layer

Tin oxide
Zinc oxide
Cadmium suflide
CIGS Cadmium Indium
Gallium diSelenide
Molybdenum

P type
bottom contact

Polyimide substrate

Figure 3.79: Cadmium Indium Gallium diSelenide solar cell (CIGS)

Table 3.5: Solar


Maximum
efficiency
Selenium, polycrystalline
0.7%
Silicon, single crystal
Silicon, single crystal PERL, ter- 25%
restrial, space
Silicon, single crystal, commer- 24%
cial terrestrial
Cypress Semiconductor, Sun- 21.5%
power, silicon single crystal
Gallium Indium Phosphide/ Gal- lium Arsenide/ Germanium, single crystal, multilayer
Advanced terrestrial version of above.
Silicon, multicrystalline
18.5%
Thin films,
Silicon, amorphous
13%

cell properties
Practical Notes
efficiency
1883, Charles Fritts
4%
1950s, first silicon solar cell
solar cars, cost=100x commercial

Cadmium telluride, polycrystalline


Copper indium arsenide diselenide, polycrystalline
Organic polymer, 100% plastic

16%

Solar cell type

14-17%

$5-$10/peak watt

19%

all contacts on cell back

32%

Preferred for space.

40.7%

Uses optical concentrator.

15.5%
5-7%

Degrades in sun light. Good indoors for calculators or cloudy


outdoors.
glass or metal substrate

18%

10%

10 inch flexible polymer web. [17]

4.5%

R&D project

3.11.

SPECIAL-PURPOSE DIODES

3.11.7

153

Varicap or varactor diodes

capacitance

A variable capacitance diode is known as a varicap diode or as a varactor. If a diode is reverse


biased, an insulating depletion region forms between the two semiconductive layers. In many diodes
the width of the depletion region may be changed by varying the reverse bias. This varies the
capacitance. This effect is accentuated in varicap diodes. The schematic symbols is shown in
Figure 3.80, one of which is packaged as common cathode dual diode.

symbol

Vcontrol

Clarge

+
voltage

Coptional

varicap diode

Figure 3.80: Varicap diode: Capacitance varies with reverse bias. This varies the frequency of a
resonant network.
If a varicap diode is part of a resonant circuit as in Figure 3.80, the frequency my be varied with
a control voltage, Vcontrol . A large capacitance, low Xc , in series with the varicap prevents Vcontrol
from being shorted out by inductor L. As long as the series capacitor is large, it has minimal effect
on the frequency of resonant circuit. Coptional may be used to set the center resonant frequency.
Vcontrol can then vary the frequency about this point. Note that the required active circuitry to
make the resonant network oscillate is not shown.
Some varicap diodes may be referred to as abrupt, hyperabrupt, or super hyper abrupt. These
refer to the change in junction capacitance with changing reverse bias as being abrupt or hyperabrupt, or super hyperabrupt. These diodes offer a relatively large change in capacitance. This
is useful when oscillators or filters are swept over a large frequency range. Varying the bias of
abrupt varicaps over the rated limits, changes capacitance by a 4:1 ratio, hyperabrupt by 10:1,
super hyperabrupt by 20:1.
Varactor diodes may be used in frequency multiplier circuits. See Practical analog semiconductor circuits, Varactor multiplier

3.11.8

Snap diode

The snap diode, also known as the step recovery diode is designed for use in high ratio frequency
multipliers up to 20 gHz. When the diode is forward biased, charge is stored in the PN junction.
This charge is drawn out as the diode is reverse biased. The diode looks like a low impedance
current source during forward bias. When reverse bias is applied it still looks like a low impedance
source until all the charge is withdrawn. It then snaps to a high impedance state causing a voltage
impulse, rich in harmonics. An applications is a comb generator, a generator of many harmonics.
Moderate power 2x and 4x multipliers are another application.

154

3.11.9

CHAPTER 3. DIODES AND RECTIFIERS

PIN diodes

A PIN diode is a fast low capacitance switching diode. Do not confuse a PIN switching diode with
a PIN photo diode (page 147). A PIN diode is manufactured like a silicon switching diode with an
intrinsic region added between the PN junction layers. This yields a thicker depletion region, the
insulating layer at the junction of a reverse biased diode. This results in lower capacitance than a
reverse biased switching diode.

top metal contact


p+ contact region
p diffusion
intrinsic region
(larger depletion
region)
n type
n+ contact region
bottom metal contact

Figure 3.81: Pin diode: Cross section aligned with schematic symbol.
PIN diodes are used in place of switching diodes in radio frequency (RF) applications, for example, a T/R switch (page 376). The 1n4007 1000 V, 1 A general purpose power diode is reported
to be useable as a PIN switching diode. The high voltage rating of this diode is achieved by the
inclusion of an intrinsic layer dividing the PN junction. This intrinsic layer makes the 1n4007 a
PIN diode. Another PIN diode application is a the antenna switch (page 376) for a direction finder
receiver.
PIN diodes serve as variable resistors when the forward bias is varied. One such application is the
voltage variable attenuator (page 376). The low capacitance characteristic of PIN diodes, extends
the frequency flat response of the attenuator to microwave frequencies.

3.11.10

IMPATT diode

The IMPact Avalanche Transit Time diode is a high power radio frequency (RF) generator operating
from 3 to 100 gHz. IMPATT diodes are fabricated from silicon, gallium arsenide, or silicon carbide.
An IMPATT diode is reverse biased above the breakdown voltage. The high doping levels produce
a thin depletion region. The resulting high electric field rapidly accelerates carriers which free other
carriers in collisions with the crystal lattice. Holes are swept into the P+ region. Electrons drift
toward the N regions. The cascading effect creates an avalanche current which increases even as
voltage across the junction decreases. The pulses of current lag the voltage peak across the junction.

3.11.

SPECIAL-PURPOSE DIODES

155

A negative resistance effect in conjunction with a resonant circuit produces oscillations at high
power levels (high for semiconductors).

N+
N-

choke
drift

N+
avalanche

P+

resonant circuit

Figure 3.82: IMPATT diode: Oscillator circuit and heavily doped P and N layers.
The resonant circuit in the schematic diagram of Figure 3.82 is the lumped circuit equivalent of
a waveguide section, where the IMPATT diode is mounted. DC reverse bias is applied through a
choke which keeps RF from being lost in the bias supply. This may be a section of waveguide known
as a bias Tee. Low power RADAR transmitters may use an IMPATT diode as a power source. They
are too noisy for use in the receiver. [20]

3.11.11

Gunn diode

Diode, gunn Gunn diode


A gunn diode is solely composed of N-type semiconductor. As such, it is not a true diode.
Figure 3.83 shows a lightly doped N layer surrounded by heavily doped N+ layers. A voltage
applied across the N-type gallium arsenide gunn diode creates a strong electric field across the
lightly doped N layer.

choke

N+

threshold

N+
resonant circuit

Figure 3.83: Gunn diode: Oscillator circuit and cross section of only N-type semiconductor diode.
As voltage is increased, conduction increases due to electrons in a low energy conduction band.
As voltage is increased beyond the threshold of approximately 1 V, electrons move from the lower
conduction band to the higher energy conduction band where they no longer contribute to conduction. In other words, as voltage increases, current decreases, a negative resistance condition. The

156

CHAPTER 3. DIODES AND RECTIFIERS

oscillation frequency is determined by the transit time of the conduction electrons, which is inversely
related to the thickness of the N layer.
The frequency may be controlled to some extent by embedding the gunn diode into a resonant
circuit. The lumped circuit equivalent shown in Figure 3.83 is actually a coaxial transmission line
or waveguide. Gallium arsenide gunn diodes are available for operation from 10 to 200 gHz at 5 to
65 mw power. Gunn diodes may also serve as amplifiers. [19] [14]

3.11.12

Shockley diode

The Shockley diodeis a 4-layer thyristor used to trigger larger thyristors. It only conducts in one
direction when triggered by a voltage exceeding the breakover voltage, about 20 V. See Thyristors,
The Shockley Diode. The bidirectional version is called a diac. See Thyristors, The DIAC.

3.11.13

Constant-current diodes

A constant-current diode, also known as a current-limiting diode, or current-regulating diode, does


exactly what its name implies: it regulates current through it to some maximum level. The constant
current diode is a two terminal version of a JFET. If we try to force more current through a constantcurrent diode than its current-regulation point, it simply fights back by dropping more voltage. If
we were to build the circuit in Figure 3.84(a) and plot diode current against diode voltage, wed get
a graph that rises at first and then levels off at the current regulation point as in Figure 3.84(b).
Rdropping

Idiode

constant-current
diode
(a)

(b)
Vdiode

Figure 3.84: Constant current diode: (a) Test circuit, (b) current vs voltage characteristic.
One application for a constant-current diode is to automatically limit current through an LED
or laser diode over a wide range of power supply voltages as in Figure ??.
Of course, the constant-current diodes regulation point should be chosen to match the LED or
laser diodes optimum forward current. This is especially important for the laser diode, not so much
for the LED, as regular LEDs tend to be more tolerant of forward current variations.
Another application is in the charging of small secondary-cell batteries, where a constant charging
current leads to predictable charging times. Of course, large secondary-cell battery banks might also
benefit from constant-current charging, but constant-current diodes tend to be very small devices,
limited to regulating currents in the milliamp range.

3.12.

OTHER DIODE TECHNOLOGIES

3.12

Other diode technologies

3.12.1

SiC diodes

157

Diodes manufactured from silicon carbide are capable of high temperature operation to 400 o C. This
could be in a high temperature environment: down hole oil well logging, gas turbine engines, auto
engines. Or, operation in a moderate environment at high power dissipation. Nuclear and space
applications are promising as SiC is 100 times more resistant to radiation compared with silicon.
SiC is a better conductor of heat than any metal. Thus, SiC is better than silicon at conducting
away heat. Breakdown voltage is several kV. SiC power devices are expected to reduce electrical
energy losses in the power industry by a factor of 100.

3.12.2

Polymer diode

Diodes based on organic chemicals have been produce using low temperature processes. Hole rich
and electron rich conductive polymers may be ink jet printed in layers. Most of the research and
development is of the organic LED (OLED). However, development of inexpensive printable organic
RFID (radio frequency identification) tags is on going. In this effort, a pentacene organic rectifier has
been operated at 50 MHz. Rectification to 800 MHz is a development goal. An inexpensive metal
insulator metal (MIM) diode acting like a back-to-back zener diode clipper has been delveloped.
Also, a tunnel diode like device has been fabricated.

3.13

SPICE models

The SPICE circuit simulation program provides for modeling diodes in circuit simulations. The
diode model is based on characterization of individual devices as described in a product data sheet
and manufacturing process characteristics not listed. Some information has been extracted from a
1N4004 data sheet in Figure 3.85.
The diode statement begins with a diode element name which must begin with d plus optional
characters. Example diode element names include: d1, d2, dtest, da, db, d101. Two node numbers
specify the connection of the anode and cathode, respectively, to other components. The node
numbers are followed by a model name, referring to a subsequent .model statement.
The model statement line begins with .model, followed by the model name matching one or
more diode statements. Next, a d indicates a diode is being modeled. The remainder of the model
statement is a list of optional diode parameters of the form ParameterName=ParameterValue. None
are used in Example below. Example2 has some parameters defined. For a list of diode parameters,
see Table 3.6.
General form:

d[name] [anode] [cathode] [model]


.model ([modelname] d [parmtr1=x] [parmtr2=y] . . .)

Example:

d1 1 2 mod1
.model mod1 d

Example2:

D2 1 2 Da1N4004
.model Da1N4004 D (IS=18.8n RS=0 BV=400 IBV=5.00u CJO=30 M=0.333 N=2)

CHAPTER 3. DIODES AND RECTIFIERS


100
CJ junction capacitance (pF)

10

1.0

0.1
0.925

IF instaneous forward current (I)

158

0.01
0.6

0.8

1.0

1.2

1.4

30

10

1
1

1.6

10

100

VR reverse voltage (V)

VF instaneous forward voltage (V)


1

Forward voltage drop VF (V)

Peak repetitive reverse voltage VRRM (V) 400

@ IF (A)

Max reverse current IR ()

Max avg rectified current IO (A)


Peak forward surge current IFSM (A)
Total capacitance CT (pF)

30
15

@ VR (V) 400

Figure 3.85: Data sheet 1N4004 excerpt, after [6].

The easiest approach to take for a SPICE model is the same as for a data sheet: consult the
manufacturers web site. Table 3.7 lists the model parameters for some selected diodes. A fallback
strategy is to build a SPICE model from those parameters listed on the data sheet. A third strategy,
not considered here, is to take measurements of an actual device. Then, calculate, compare and
adjust the SPICE parameters to the measurements.
If diode parameters are not specified as in Example model above, the parameters take on the
default values listed in Table 3.6 and Table 3.7. These defaults model integrated circuit diodes.
These are certainly adequate for preliminary work with discrete devices For more critical work, use
SPICE models supplied by the manufacturer [5], SPICE vendors, and other sources. [16]
Otherwise, derive some of the parameters from the data sheet. First select a value for spice
parameter N between 1 and 2. It is required for the diode equation (n). Massobrio [1] pp 9,
recommends .. n, the emission coefficient is usually about 2. In Table 3.7, we see that power
rectifiers 1N3891 (12 A), and 10A04 (10 A) both use about 2. The first four in the table are not
relevant because they are schottky, schottky, germanium, and silicon small signal, respectively. The
saturation current, IS, is derived from the diode equation, a value of (VD , ID ) on the graph in
Figure 3.85, and N=2 (n in the diode equation).
ID = IS (eVD /nVT 1)
VT = 26 mV at 25o C

n = 2.0

VD = 0.925 V at 1 A from graph

3.13.

SPICE MODELS

Symbol
IS

Name
IS

RS

RS

n
D
CD (0)
0
m
-

N
TT
CJO
VJ
M
-

Eg
pi
kf
af
FC

EG
XTI
KF
AF
FC

BV
IBV

BV
IBV

159

Table 3.6: Diode SPICE parameters


Parameter
Saturation current (diode equation)
Parsitic resistance (series resistance)
Emission coefficient, 1 to 2
Transit time
Zero-bias junction capacitance
Junction potential
Junction grading coefficient
0.33 for linearly graded junction
0.5 for abrupt junction
Activation energy:
Si: 1.11
Ge: 0.67
Schottky: 0.69
IS temperature exponent
pn junction: 3.0
Schottky: 2.0
Flicker noise coefficient
Flicker noise exponent
Forward bias depletion capacitance coefficient
Reverse breakdown voltage
Reverse breakdown current

Table 3.7: SPICE parameters


Part
IS
RS
Default
1E-14 0
1N5711 sk
315n
2.8
1N5712 sk
680p
12
1N34 Ge
200p
84m
1N4148
35p
64m
1N3891
63n
9.6m
10A04 10A 844n
2.06m
1N4004 1A 76.9n 42.2m
1N4004
18.8n data sheet

Units
A

Default
1E-14

s
F
V
-

1
0
0
1
0.5
-

eV
-

1.11
3.0
0
1
0.5

V
A

1E-3

for selected diodes; sk=schottky Ge=germanium; else silicon.


N
TT
CJO M
VJ EG XTI BV IBV
1
0
0
0.5
1
1.11 3

1m
2.03
1.44n 2.00p 0.333 0.69 2
70
10u
1.003 50p
1.0p
0.5
0.6 0.69 2
20
2.19
144n 4.82p 0.333 0.75 0.67 60
15u
1.24
5.0n
4.0p
0.285 0.6 75
2
110n 114p 0.255 0.6 250 2.06
4.32u 277p 0.333 400 10u
1.45
4.32u 39.8p 0.333 400 5u
2
30p
0.333 400 5u

160

CHAPTER 3. DIODES AND RECTIFIERS


1 A = IS (e(0.925V )/(2)(26mV ) 1)
IS =

18.8E-9

The numerical values of IS=18.8n and N=2 are entered in last line of Table 3.7 for comparison
to the manufacturers model for 1N4004, which is considerably different. RS defaults to 0 for now.
It will be estimated later. The important DC static parameters are N, IS, and RS.
Rashid [15] suggests that TT, D , the transit time, be approximated from the reverse recovery
stored charge QRR , a data sheet parameter (not available on our data sheet) and IF , forward current.
ID = IS (eVD /nVT 1)
D =

QRR /IF

We take the TT=0 default for lack of QRR . Though it would be reasonable to take TT for a
similar rectifier like the 10A04 at 4.32u. The 1N3891 TT is not a valid choice because it is a fast
recovery rectifier. CJO, the zero bias junction capacitance is estimated from the V R vs CJ graph
in Figure 3.85. The capacitance at the nearest to zero voltage on the graph is 30 pF at 1 V. If
simulating high speed transient response, as in switching regulator power supplies, TT and CJO
parameters must be provided.
The junction grading coefficient M is related to the doping profile of the junction. This is not a
data sheet item. The default is 0.5 for an abrupt junction. We opt for M=0.333 corresponding to a
linearly graded junction. The power rectifiers in Table 3.7 use lower values for M than 0.5.
We take the default values for VJ and EG. Many more diodes use VJ=0.6 than shown in Table 3.7.
However the 10A04 rectifier uses the default, which we use for our 1N4004 model (Da1N4001 in
Table 3.6). Use the default EG=1.11 for silicon diodes and rectifiers. Table 3.6 lists values for
schottky and germanium diodes. Take the XTI=3, the default IS temperature coefficient for silicon
devices. See Table 3.6 for XTI for schottky diodes.
The abbreviated data sheet, Figure 3.85, lists IR = 5 A @ VR = 400 V, corresponding to
IBV=5u and BV=400 respectively. The 1n4004 SPICE parameters derived from the data sheet are
listed in the last line of Table 3.7 for comparison to the manufacturers model listed above it. BV
is only necessary if the simulation exceeds the reverse breakdown voltage of the diode, as is the
case for zener diodes. IBV, reverse breakdown current, is frequently omitted, but may be entered if
provided with BV.
Figure 3.86 shows a circuit to compare the manufacturers model, the model derived from the
datasheet, and the default model using default parameters. The three dummy 0 V sources are
necessary for diode current measurement. The 1 V source is swept from 0 to 1.4 V in 0.2 mV
steps. See .DC statement in the netlist in Table 3.8. DI1N4004 is the manufacturers diode model,
Da1N4004 is our derived diode model.
We compare the three models in Figure 3.87. and to the datasheet graph data in Table 3.9. VD
is the diode voltage versus the diode currents for the manufacturers model, our calculated datasheet
model and the default diode model. The last column 1N4004 graph is from the datasheet voltage
versus current curve in Figure 3.85 which we attempt to match. Comparison of the currents for
the three model to the last column shows that the default model is good at low currents, the
manufacturers model is good at high currents, and our calculated datasheet model is best of all up

3.13.

SPICE MODELS

161

D1

+
1V
+
0V
0

D2

+
0V

D3

+
0V

Figure 3.86: SPICE circuit for comparison of manufacturer model (D1), calculated datasheet model
(D2), and default model (D3).

Table 3.8: SPICE netlist parameters: (D1) DI1N4004 manufacturers model, (D2) Da1N40004
datasheet derived, (D3) default diode model.
*SPICE circuit <03468.eps> from XCircuit v3.20
D1 1 5 DI1N4004
V1 5 0 0
D2 1 3 Da1N4004
V2 3 0 0
D3 1 4 Default
V3 4 0 0
V4 1 0 1
.DC V4 0 1400mV 0.2m
.model Da1N4004 D (IS=18.8n RS=0
BV=400 IBV=5.00u CJO=30
+M=0.333
N=2.0 TT=0)
.MODEL DI1N4004 D (IS=76.9n RS=42.0m BV=400 IBV=5.00u CJO=39.8p
+M=0.333 N=1.45 TT=4.32u)
.MODEL Default D
.end

162

CHAPTER 3. DIODES AND RECTIFIERS

to 1 A. Agreement is almost perfect at 1 A because the IS calculation is based on diode voltage at


1 A. Our model grossly over states current above 1 A.

Figure 3.87: First trial of manufacturer model, calculated datasheet model, and default model.

Table 3.9: Comparison of manufacturer model, calculated datasheet model, and default model to
1N4004 datasheet graph of V vs I.

index
3500
4001
4500
4625
5000
5500
6000
6500
7000

VD
7.000000e-01
8.002000e-01
9.000000e-01
9.250000e-01
1.000000e-00
1.100000e+00
1.200000e+00
1.300000e+00
1.400000e+00

model
manufacturer
1.612924e+00
3.346832e+00
5.310740e+00
5.823654e+00
7.395953e+00
9.548779e+00
1.174489e+01
1.397087e+01
1.621861e+01

model
datasheet
1.416211e-02
9.825960e-02
6.764928e-01
1.096870e+00
4.675526e+00
3.231452e+01
2.233392e+02
1.543591e+03
1.066840e+04

model
1N4004
default
graph
5.674683e-03 0.01
2.731709e-01 0.13
1.294824e+01 0.7
3.404037e+01 1.0
6.185078e+02 2.0
2.954471e+04 3.3
1.411283e+06 5.3
6.741379e+07 8.0
3.220203e+09 12.

The solution is to increase RS from the default RS=0. Changing RS from 0 to 8m in the datasheet
model causes the curve to intersect 10 A (not shown) at the same voltage as the manufacturers
model. Increasing RS to 28.6m shifts the curve further to the right as shown in Figure 3.88. This
has the effect of more closely matching our datasheet model to the datasheet graph (Figure 3.85).
Table 3.10 shows that the current 1.224470e+01 A at 1.4 V matches the graph at 12 A. However,
the current at 0.925 V has degraded from 1.096870e+00 above to 7.318536e-01.
Suggested reader exercise: decrease N so that the current at VD=0.925 V is restored to 1 A.

3.13.

SPICE MODELS

163

Figure 3.88: Second trial to improve calculated datasheet model compared with manufacturer model
and default model.

Table 3.10: Changing Da1N4004 model statement RS=0 to RS=28.6m decreases the current at
VD=1.4 V to 12.2 A.
.model Da1N4004 D (IS=18.8n RS=28.6m BV=400 IBV=5.00u CJO=30
+M=0.333
N=2.0 TT=0)
model
model
1N4001
index
VD
manufacturer
datasheet
graph
3505
7.010000e-01
1.628276e+00
1.432463e-02
0.01
4000
8.000000e-01
3.343072e+00
9.297594e-02
0.13
4500
9.000000e-01
5.310740e+00
5.102139e-01
0.7
4625
9.250000e-01
5.823654e+00
7.318536e-01
1.0
5000
1.000000e-00
7.395953e+00
1.763520e+00
2.0
5500
1.100000e+00
9.548779e+00
3.848553e+00
3.3
6000
1.200000e+00
1.174489e+01
6.419621e+00
5.3
6500
1.300000e+00
1.397087e+01
9.254581e+00
8.0
7000
1.400000e+00
1.621861e+01
1.224470e+01
12.

164

CHAPTER 3. DIODES AND RECTIFIERS

This may increase the current (12.2 A) at VD=1.4 V requiring an increase of RS to decrease current
to 12 A.
Zener diode: There are two approaches to modeling a zener diode: set the BV parameter to
the zener voltage in the model statement, or model the zener with a subcircuit containing a diode
clamper set to the zener voltage. An example of the first approach sets the breakdown voltage BV
to 15 for the 1n4469 15 V zener diode model (IBV optional):
.model D1N4469 D ( BV=15 IBV=17m )
The second approach models the zener with a subcircuit. Clamper D1 and VZ in Figure ??
models the 15 V reverse breakdown voltage of a 1N4477A zener diode. Diode DR accounts for the
forward conduction of the zener in the subcircuit.

D1

+ 13.7 V
3

.SUBCKT DI-1N4744A 1
2
* Terminals A K
D1 1 2 DF
DZ 3 1 DR
VZ 2 3 13.7
.MODEL DF D (
IS=27.5p RS=0.620
N=1.10
+ CJO=78.3p VJ=1.00
M=0.330 TT=50.1n )
.MODEL DR D (
IS=5.49f RS=0.804
N=1.77 )
.ENDS

Figure 3.89: Zener diode subcircuit uses clamper (D1 and VZ) to model zener.
Tunnel diode: A tunnel diode may be modeled by a pair of field effect transistors (JFET) in
a SPICE subcircuit. [11] An oscillator circuit is also shown in this reference.
Gunn diode: A Gunn diode may also be modeled by a pair of JFETs. [12] This reference
shows a microwave relaxation oscillator.
REVIEW:
Diodes are described in SPICE by a diode component statement referring to .model statement. The .model statement contains parameters describing the diode. If parameters are not
provided, the model takes on default values.
Static DC parameters include N, IS, and RS. Reverse breakdown parameters: BV, IBV.
Accurate dynamic timing requires TT and CJO parameters
Models provided by the manufacturer are highly recommended.

BIBLIOGRAPHY

165

Contributors
Contributors to this chapter are listed in chronological order of their contributions, from most recent
to first. See Appendix 2 (Contributor List) for dates and contact information.
Jered Wierzbicki (December 2002): Pointed out error in diode equation Boltzmanns constant shown incorrectly.

Bibliography
[1] Paolo Antognetti, Giuseppe Massobrio Semiconductor Device Modeling with SPICE, ISBN
0-07-002107-4, 1988
[2] ATCO
Newsletter,
Volume
14
http://www.atco.tv/homepage/vol14 1.pdf

No.

1,

January

1997

at

[3] D.A. Brunner, et al,, A Cockcroft-Walton Base for the FEU84-3 Photomultiplier Tube,
Department of Physics, Indiana University, Bloomington, Indiana 47405 January 1998, at
http://dustbunny.physics.indiana.edu/~paul/cwbase/
[4] Brenton Burnet, The Basic Physics and Design of III-V Multijunction Solar, NREL, at
photochemistry.epfl.ch/EDEY/NREL.pdf
[5] Diodes Incorporated http://www.diodes.com/products/spicemodels/index.php
[6] Diodes
Incorporated,
1N4001/L
1N4007/l,
http://www.diodes.com/datasheets/ds28002.pdf
[7] Christiana
Honsberg,
Stuart
Bowden,
http://www.udel.edu/igert/pvcdrom/

1.0A

Photovoltaics

rectifier,

at

CDROM,

at

[8] R. R. King, et. al., 40% efficient metamorphic GaInP/GaInAs/Ge multijunction solar cells, Applied Physics Letters, 90, 183516 (2007) , at http://scitation.aip.org
/getabs/servlet/GetabsServlet?prog=normal&id=APPLAB000090000018183516000001
&idtype=cvips&gifs=yes
[9] Chin-Leong Lim, Lim Yeam Chng, Goh Swee Chye, Diode Quad Is Foundation For PIN Diode Attenuator,
Microwaves & RF, May 2006,
at
http://www.mwrf.com/Articles/Index.cfm?Ad=1&ArticleID=12523
[10] Kim W Mitchell, Method of making a thin film cadmium telluride solar cell, United States
Patent 4734381,http://www.freepatentsonline.com/4734381.html
[11] Karl H. Muller RF/Microwave Analysis Intusoft Newsletter #51, November 1997, at
http://www.intusoft.com/nlhtm/nl51.htm
[12] A Gunn Diode Relaxation Oscillator, Intusoft Newsletter #52, February 1998, at
http://www.intusoft.com/nlhtm/nl52.htm
[13] OAK
Solar.,
Technical
LEDs
LED
http://www.oksolar.com/led/led color chart.htm

color

chart,

at

166

BIBLIOGRAPHY

[14] Ian Poole, Summary of the Gunn Diode, at http://www.radio-electronics.com/


nfo/data/semicond/gunndiode/gunndiode.php
[15] Muhammad H. Rashid, SPICE for Power Electronics and Electric Power, ISBN 0-13-0304204, 1993
[16] SPICE
model
index,
V2.16
30-Nov-05,
~paul.hills/Circuits/Spice/ModelIndex.html

at

http://homepages.which.net/

[17] Neil Thomas, Advancing CIGS Solar Cell Manufacturing Technology, April 6, 2007 at
http://www.renewableenergyaccess.com/rea/news/story?id=48033&src=rss
[18] P.J. Verlinden,
Sinton,
K. Wickham,
R.M. Swanson Crane,
BACKSIDECONTACT SILICON SOLAR CELLS WITH IMPROVED EFFICIENCY. at
http://www.sunpowercorp.com/techpapers/EPSEC97.pdf
[19] Christian Wolff, Radar Principles, Radar components,
http://www.radartutorial.eu/17.bauteile/bt12.en.htm

Gunn

diodes

at

at

[20] L. Yuan, M. R. Melloch, J. A. Cooper, K. J. Webb,Silicon Carbide IMPATT Oscillators


for High-Power Microwave and Millimeter-Wave Generation, IEEE/Cornell Conference on
Advanced Concepts in High Speed Semiconductor Devices and Circuits, Ithaca, NY, August 79, 2000. at http://www.ecn.purdue.edu/WBG/Device Research/IMPATT Diodes/Index.html
[21] Alan Seabaugh, Zhaoming HU, Qingmin LIU, David Rink, Jinli Wang, Silicon Based Tunnel
Diodes and Integrated Circuits, at http://www.nd.edu/~nano/0a1003QFDpaper v1.pdf
[22] S. M. Sze, G. Gibbons, Avalanche breakdown voltages of abrupt and linearly graded p-n
junctions in Ge, Si, GaAs, and Ga P, Appl. Phys. Lett., 8, 111 (1966).
[23] Lisa Zyga, 40% efficient solar cells to be used for solar electricity, PhysOrgForum, at
http://www.physorg.com/news99904887.html

Chapter 4

BIPOLAR JUNCTION
TRANSISTORS
Contents
4.1
4.2
4.3
4.4
4.5
4.6
4.7
4.8
4.9
4.10
4.11
4.12
4.13
4.14

Introduction . . . . . . . . . . . .
The transistor as a switch . . . .
Meter check of a transistor . . .
Active mode operation . . . . . .
The common-emitter amplifier .
The common-collector amplifier
The common-base amplifier . . .
Biasing techniques . . . . . . . .
Input and output coupling . . .
Feedback . . . . . . . . . . . . . .
Amplifier impedances . . . . . .
Current mirrors . . . . . . . . . .
Transistor ratings and packages
BJT quirks . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

167
170
173
178
186
201
209
215
228
235
242
243
246
247

*** INCOMPLETE ***

4.1

Introduction

The invention of the bipolar transistor in 1948 ushered in a revolution in electronics. Technical
feats previously requiring relatively large, mechanically fragile, power-hungry vacuum tubes were
suddenly achievable with tiny, mechanically rugged, power-thrifty specks of crystalline silicon. This
revolution made possible the design and manufacture of lightweight, inexpensive electronic devices
that we now take for granted. Understanding how transistors function is of paramount importance
to anyone interested in understanding modern electronics.
167

168

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

My intent here is to focus as exclusively as possible on the practical function and application of
bipolar transistors, rather than to explore the quantum world of semiconductor theory. Discussions
of holes and electrons are better left to another chapter in my opinion. Here I want to explore how
to use these components, not analyze their intimate internal details. I dont mean to downplay the
importance of understanding semiconductor physics, but sometimes an intense focus on solid-state
physics detracts from understanding these devices functions on a component level. In taking this
approach, however, I assume that the reader possesses a certain minimum knowledge of semiconductors: the difference between P and N doped semiconductors, the functional characteristics of a
PN (diode) junction, and the meanings of the terms reverse biased and forward biased. If these
concepts are unclear to you, it is best to refer to earlier chapters in this book before proceeding with
this one.
A bipolar transistor consists of a three-layer sandwich of doped (extrinsic) semiconductor
materials, either P-N-P or N-P-N. Each layer forming the transistor has a specific name, and each
layer is provided with a wire contact for connection to a circuit. Shown here are schematic symbols
and physical diagrams of these two transistor types:

PNP transistor
collector
collector
base

base

P
N
P

emitter
emitter
schematic symbol

physical diagram
NPN transistor
collector

collector
base

base

N
P
N

emitter
emitter
schematic symbol

physical diagram

4.1.

INTRODUCTION

169

The only functional difference between a PNP transistor and an NPN transistor is the proper
biasing (polarity) of the junctions when operating. For any given state of operation, the current
directions and voltage polarities for each type of transistor are exactly opposite each other.
Bipolar transistors work as current-controlled current regulators. In other words, they restrict
the amount of current that can go through them according to a smaller, controlling current. The
main current that is controlled goes from collector to emitter, or from emitter to collector, depending
on the type of transistor it is (PNP or NPN, respectively). The small current that controls the main
current goes from base to emitter, or from emitter to base, once again depending on the type of
transistor it is (PNP or NPN, respectively). According to the confusing standards of semiconductor
symbology, the arrow always points against the direction of electron flow:

C
E

= small, controlling current


= large, controlled current
Bipolar transistors are called bi polar because the main flow of electrons through them takes place
in two types of semiconductor material: P and N, as the main current goes from emitter to collector
(or vice versa). In other words, two types of charge carriers electrons and holes comprise this
main current through the transistor.
As you can see, the controlling current and the controlled current always mesh together through
the emitter wire, and their electrons always flow against the direction of the transistors arrow. This
is the first and foremost rule in the use of transistors: all currents must be going in the proper
directions for the device to work as a current regulator. The small, controlling current is usually
referred to simply as the base current because it is the only current that goes through the base
wire of the transistor. Conversely, the large, controlled current is referred to as the collector current
because it is the only current that goes through the collector wire. The emitter current is the sum
of the base and collector currents, in compliance with Kirchhoffs Current Law.
If there is no current through the base of the transistor, it shuts off like an open switch and
prevents current through the collector. If there is a base current, then the transistor turns on like a
closed switch and allows a proportional amount of current through the collector. Collector current
is primarily limited by the base current, regardless of the amount of voltage available to push it.
The next section will explore in more detail the use of bipolar transistors as switching elements.
REVIEW:
Bipolar transistors are so named because the controlled current must go through two types

170

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS


of semiconductor material: P and N. The current consists of both electron and hole flow, in
different parts of the transistor.
Bipolar transistors consist of either a P-N-P or an N-P-N semiconductor sandwich structure.
The three leads of a bipolar transistor are called the Emitter, Base, and Collector.
Transistors function as current regulators by allowing a small current to control a larger current.
The amount of current allowed between collector and emitter is primarily determined by the
amount of current moving between base and emitter.
In order for a transistor to properly function as a current regulator, the controlling (base)
current and the controlled (collector) currents must be going in the proper directions: meshing
additively at the emitter and going against the emitter arrow symbol.

4.2

The transistor as a switch

Because a transistors collector current is proportionally limited by its base current, it can be used as
a sort of current-controlled switch. A relatively small flow of electrons sent through the base of the
transistor has the ability to exert control over a much larger flow of electrons through the collector.
Suppose we had a lamp that we wanted to turn on and off by means of a switch. Such a circuit
would be extremely simple:

switch

For the sake of illustration, lets insert a transistor in place of the switch to show how it can
control the flow of electrons through the lamp. Remember that the controlled current through a
transistor must go between collector and emitter. Since its the current through the lamp that we
want to control, we must position the collector and emitter of our transistor where the two contacts
of the switch are now. We must also make sure that the lamps current will move against the
direction of the emitter arrow symbol to ensure that the transistors junction bias will be correct:

NPN
transistor

In this example I happened to choose an NPN transistor. A PNP transistor could also have been
chosen for the job, and its application would look like this:

4.2.

THE TRANSISTOR AS A SWITCH

171

PNP
transistor

The choice between NPN and PNP is really arbitrary. All that matters is that the proper current
directions are maintained for the sake of correct junction biasing (electron flow going against the
transistor symbols arrow).
Going back to the NPN transistor in our example circuit, we are faced with the need to add
something more so that we can have base current. Without a connection to the base wire of the
transistor, base current will be zero, and the transistor cannot turn on, resulting in a lamp that is
always off. Remember that for an NPN transistor, base current must consist of electrons flowing
from emitter to base (against the emitter arrow symbol, just like the lamp current). Perhaps the
simplest thing to do would be to connect a switch between the base and collector wires of the
transistor like this:

switch

If the switch is open, the base wire of the transistor will be left floating (not connected to
anything) and there will be no current through it. In this state, the transistor is said to be cutoff.
If the switch is closed, however, electrons will be able to flow from the emitter through to the base
of the transistor, through the switch and up to the left side of the lamp, back to the positive side of
the battery. This base current will enable a much larger flow of electrons from the emitter through
to the collector, thus lighting up the lamp. In this state of maximum circuit current, the transistor
is said to be saturated.

switch

Of course, it may seem pointless to use a transistor in this capacity to control the lamp. After
all, were still using a switch in the circuit, arent we? If were still using a switch to control the
lamp if only indirectly then whats the point of having a transistor to control the current? Why
not just go back to our original circuit and use the switch directly to control the lamp current?
There are a couple of points to be made here, actually. First is the fact that when used in
this manner, the switch contacts need only handle what little base current is necessary to turn
the transistor on, while the transistor itself handles the majority of the lamps current. This may

172

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

be an important advantage if the switch has a low current rating: a small switch may be used
to control a relatively high-current load. Perhaps more importantly, though, is the fact that the
current-controlling behavior of the transistor enables us to use something completely different to
turn the lamp on or off. Consider this example, where a solar cell is used to control the transistor,
which in turn controls the lamp:

solar
cell
Or, we could use a thermocouple to provide the necessary base current to turn the transistor on:

thermocouple
+
source of
heat
Even a microphone of sufficient voltage and current output could be used to turn the transistor
on, provided its output is rectified from AC to DC so that the emitter-base PN junction within the
transistor will always be forward-biased:

microphone

source of
sound

The point should be quite apparent by now: any sufficient source of DC current may be used to
turn the transistor on, and that source of current need only be a fraction of the amount of current
needed to energize the lamp. Here we see the transistor functioning not only as a switch, but as
a true amplifier : using a relatively low-power signal to control a relatively large amount of power.
Please note that the actual power for lighting up the lamp comes from the battery to the right of
the schematic. It is not as though the small signal current from the solar cell, thermocouple, or

4.3.

METER CHECK OF A TRANSISTOR

173

microphone is being magically transformed into a greater amount of power. Rather, those small
power sources are simply controlling the batterys power to light up the lamp.
REVIEW:
Transistors may be used as switching elements to control DC power to a load. The switched
(controlled) current goes between emitter and collector, while the controlling current goes
between emitter and base.
When a transistor has zero current through it, it is said to be in a state of cutoff (fully
nonconducting).
When a transistor has maximum current through it, it is said to be in a state of saturation
(fully conducting).

4.3

Meter check of a transistor

Bipolar transistors are constructed of a three-layer semiconductor sandwich, either PNP or NPN.
As such, they register as two diodes connected back-to-back when tested with a multimeters resistance or diode check functions:

A
OFF

PNP transistor
collector
+

COM

collector
base

base
emitter

P
N
P
+
emitter

A
OFF

COM

Both meters show continuity (low


resistance) through collector-base
and emitter-base PN junctions.

Here Im assuming the use of a multimeter with only a single continuity range (resistance)
function to check the PN junctions. Some multimeters are equipped with two separate continuity
check functions: resistance and diode check, each with its own purpose. If your meter has a

174

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

designated diode check function, use that rather than the resistance range, and the meter will
display the actual forward voltage of the PN junction and not just whether or not it conducts current.

A
A

V
OFF

PNP transistor
collector
-

COM

collector
base

base

+
+

P
N
P

emitter
emitter
V

A
OFF

COM

Both meters show non-continuity


(high resistance) through collectorbase and emitter-base PN junctions.

Meter readings will be exactly opposite, of course, for an NPN transistor, with both PN junctions
facing the other way. If a multimeter with a diode check function is used in this test, it will be found
that the emitter-base junction possesses a slightly greater forward voltage drop than the collectorbase junction. This forward voltage difference is due to the disparity in doping concentration between
the emitter and collector regions of the transistor: the emitter is a much more heavily doped piece
of semiconductor material than the collector, causing its junction with the base to produce a higher
forward voltage drop.
Knowing this, it becomes possible to determine which wire is which on an unmarked transistor.
This is important because transistor packaging, unfortunately, is not standardized. All bipolar
transistors have three wires, of course, but the positions of the three wires on the actual physical
package are not arranged in any universal, standardized order.
Suppose a technician finds a bipolar transistor and proceeds to measure continuity with a multimeter set in the diode check mode. Measuring between pairs of wires and recording the values
displayed by the meter, the technician obtains the following data:

4.3.

METER CHECK OF A TRANSISTOR

175

Unknown bipolar transistor

Which wires are emitter,


base, and collector?

1
2

Meter touching wire 1 (+) and 2 (-): OL


Meter touching wire 1 (-) and 2 (+): OL
Meter touching wire 1 (+) and 3 (-): 0.655 volts
Meter touching wire 1 (-) and 3 (+): OL
Meter touching wire 2 (+) and 3 (-): 0.621 volts
Meter touching wire 2 (-) and 3 (+): OL
The only combinations of test points giving conducting meter readings are wires 1 and 3 (red
test lead on 1 and black test lead on 3), and wires 2 and 3 (red test lead on 2 and black test lead on
3). These two readings must indicate forward biasing of the emitter-to-base junction (0.655 volts)
and the collector-to-base junction (0.621 volts).
Now we look for the one wire common to both sets of conductive readings. It must be the base
connection of the transistor, because the base is the only layer of the three-layer device common to
both sets of PN junctions (emitter-base and collector-base). In this example, that wire is number
3, being common to both the 1-3 and the 2-3 test point combinations. In both those sets of meter
readings, the black (-) meter test lead was touching wire 3, which tells us that the base of this
transistor is made of N-type semiconductor material (black = negative). Thus, the transistor is an
PNP type with base on wire 3, emitter on wire 1 and collector on wire 2:

1
Emitter

2
3
Collector Base

176

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

Please note that the base wire in this example is not the middle lead of the transistor, as one
might expect from the three-layer sandwich model of a bipolar transistor. This is quite often
the case, and tends to confuse new students of electronics. The only way to be sure which lead is
which is by a meter check, or by referencing the manufacturers data sheet documentation on that
particular part number of transistor.
Knowing that a bipolar transistor behaves as two back-to-back diodes when tested with a conductivity meter is helpful for identifying an unknown transistor purely by meter readings. It is also
helpful for a quick functional check of the transistor. If the technician were to measure continuity
in any more than two or any less than two of the six test lead combinations, he or she would immediately know that the transistor was defective (or else that it wasnt a bipolar transistor but rather
something else a distinct possibility if no part numbers can be referenced for sure identification!).
However, the two diode model of the transistor fails to explain how or why it acts as an amplifying
device.
To better illustrate this paradox, lets examine one of the transistor switch circuits using the
physical diagram rather than the schematic symbol to represent the transistor. This way the two
PN junctions will be easier to see:

N
P
N

collector
base
emitter

solar
cell
A grey-colored diagonal arrow shows the direction of electron flow through the emitter-base
junction. This part makes sense, since the electrons are flowing from the N-type emitter to the Ptype base: the junction is obviously forward-biased. However, the base-collector junction is another
matter entirely. Notice how the grey-colored thick arrow is pointing in the direction of electron
flow (upwards) from base to collector. With the base made of P-type material and the collector
of N-type material, this direction of electron flow is clearly backwards to the direction normally
associated with a PN junction! A normal PN junction wouldnt permit this backward direction of
flow, at least not without offering significant opposition. However, when the transistor is saturated,
there is very little opposition to electrons all the way from emitter to collector, as evidenced by the
lamps illumination!
Clearly then, something is going on here that defies the simple two-diode explanatory model
of the bipolar transistor. When I was first learning about transistor operation, I tried to construct
my own transistor from two back-to-back diodes, like this:

4.3.

METER CHECK OF A TRANSISTOR

177

no light!

no current!

solar
cell
Back-to-back diodes dont act
like a transistor!
My circuit didnt work, and I was mystified. However useful the two diode description of a
transistor might be for testing purposes, it doesnt explain how a transistor can behave as a controlled
switch.
What happens in a transistor is this: the reverse bias of the base-collector junction prevents
collector current when the transistor is in cutoff mode (that is, when there is no base current).
However, when the base-emitter junction is forward biased by the controlling signal, the normallyblocking action of the base-collector junction is overridden and current is permitted through the
collector, despite the fact that electrons are going the wrong way through that PN junction. This
action is dependent on the quantum physics of semiconductor junctions, and can only take place
when the two junctions are properly spaced and the doping concentrations of the three layers are
properly proportioned. Two diodes wired in series fail to meet these criteria, and so the top diode
can never turn on when it is reversed biased, no matter how much current goes through the bottom
diode in the base wire loop.
That doping concentrations play a crucial part in the special abilities of the transistor is further
evidenced by the fact that collector and emitter are not interchangeable. If the transistor is merely
viewed as two back-to-back PN junctions, or merely as a plain N-P-N or P-N-P sandwich of materials,
it may seem as though either end of the transistor could serve as collector or emitter. This, however,
is not true. If connected backwards in a circuit, a base-collector current will fail to control current
between collector and emitter. Despite the fact that both the emitter and collector layers of a bipolar
transistor are of the same doping type (either N or P), they are definitely not identical!
So, current through the emitter-base junction allows current through the reverse-biased basecollector junction. The action of base current can be thought of as opening a gate for current
through the collector. More specifically, any given amount of emitter-to-base current permits a
limited amount of base-to-collector current. For every electron that passes through the emitter-base
junction and on through the base wire, there is allowed a certain, restricted number of electrons to
pass through the base-collector junction and no more.
In the next section, this current-limiting behavior of the transistor will be investigated in more
detail.
REVIEW:

178

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS


Tested with a multimeter in the resistance or diode check modes, a transistor behaves like
two back-to-back PN (diode) junctions.

The emitter-base PN junction has a slightly greater forward voltage drop than the collectorbase PN junction, due to more concentrated doping of the emitter semiconductor layer.

The reverse-biased base-collector junction normally blocks any current from going through the
transistor between emitter and collector. However, that junction begins to conduct if current
is drawn through the base wire. Base current can be thought of as opening a gate for a
certain, limited amount of current through the collector.

4.4

Active mode operation

When a transistor is in the fully-off state (like an open switch), it is said to be cutoff. Conversely,
when it is fully conductive between emitter and collector (passing as much current through the
collector as the collector power supply and load will allow), it is said to be saturated. These are the
two modes of operation explored thus far in using the transistor as a switch.
However, bipolar transistors dont have to be restricted to these two extreme modes of operation.
As we learned in the previous section, base current opens a gate for a limited amount of current
through the collector. If this limit for the controlled current is greater than zero but less than the
maximum allowed by the power supply and load circuit, the transistor will throttle the collector
current in a mode somewhere between cutoff and saturation. This mode of operation is called the
active mode.
An automotive analogy for transistor operation is as follows: cutoff is the condition where there
is no motive force generated by the mechanical parts of the car to make it move. In cutoff mode, the
brake is engaged (zero base current), preventing motion (collector current). Active mode is when
the automobile is cruising at a constant, controlled speed (constant, controlled collector current) as
dictated by the driver. Saturation is when the automobile is driving up a steep hill that prevents it
from going as fast as the driver would wish. In other words, a saturated automobile is one where
the accelerator pedal is pushed all the way down (base current calling for more collector current
than can be provided by the power supply/load circuit).
Ill set up a circuit for SPICE simulation to demonstrate what happens when a transistor is in
its active mode of operation:

4.4.

ACTIVE MODE OPERATION

179

Vammeter
2

3
0V

V1

Q1

Current I1
source
0

Q is the standard letter designation for a transistor in a schematic diagram, just as R is for
resistor and C is for capacitor. In this circuit, we have an NPN transistor powered by a battery
(V1 ) and controlled by current through a current source (I1 ). A current source is a device that
outputs a specific amount of current, generating as much or as little voltage as necessary across
its terminals to ensure that exact amount of current through it. Current sources are notoriously
difficult to find in nature (unlike voltage sources, which by contrast attempt to maintain a constant
voltage, outputting as much or as little current in the fulfillment of that task), but can be simulated
with a small collection of electronic components. As we are about to see, transistors themselves tend
to mimic the constant-current behavior of a current source in their ability to regulate current at a
fixed value.
In the SPICE simulation, Ill set the current source at a constant value of 20 A, then vary the
voltage source (V1 ) over a range of 0 to 2 volts and monitor how much current goes through it.
The dummy battery (Vammeter ) with its output of 0 volts serves merely to provide SPICE with
a circuit element for current measurement.
bipolar transistor simulation
i1 0 1 dc 20u
q1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc
.model mod1 npn
.dc v1 0 2 0.05
.plot dc i(vammeter)
.end
type
is
bf
nf
br
nr

npn
1.00E-16
100.000
1.000
1.000
1.000

180

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

The constant base current of 20 A sets a collector current limit of 2 mA, exactly 100 times as
much. Notice how flat the curve is for collector current over the range of battery voltage from 0
to 2 volts. The only exception to this featureless plot is at the very beginning, where the battery
increases from 0 volts to 0.25 volts. There, the collector current increases rapidly from 0 amps to
its limit of 2 mA.
Lets see what happens if we vary the battery voltage over a wider range, this time from 0 to 50
volts. Well keep the base current steady at 20 A:

bipolar transistor simulation


i1 0 1 dc 20u
q1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc
.model mod1 npn
.dc v1 0 50 2
.plot dc i(vammeter)
.end

type
is
bf
nf
br
nr

npn
1.00E-16
100.000
1.000
1.000
1.000

4.4.

ACTIVE MODE OPERATION

181

Same result! The collector current holds absolutely steady at 2 mA despite the fact that the
battery (v1) voltage varies all the way from 0 to 50 volts. It would appear from our simulation
that collector-to-emitter voltage has little effect over collector current, except at very low levels (just
above 0 volts). The transistor is acting as a current regulator, allowing exactly 2 mA through the
collector and no more.
Now lets see what happens if we increase the controlling (I1 ) current from 20 A to 75 A, once
again sweeping the battery (V1 ) voltage from 0 to 50 volts and graphing the collector current:

bipolar transistor simulation


i1 0 1 dc 75u
q1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc
.model mod1 npn
.dc v1 0 50 2
.plot dc i(vammeter)
.end

type
is
bf
nf
br
nr

npn
1.00E-16
100.000
1.000
1.000
1.000

182

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

Not surprisingly, SPICE gives us a similar plot: a flat line, holding steady this time at 7.5 mA
exactly 100 times the base current over the range of battery voltages from just above 0 volts to
50 volts. It appears that the base current is the deciding factor for collector current, the V 1 battery
voltage being irrelevant so long as its above a certain minimum level.
This voltage/current relationship is entirely different from what were used to seeing across a
resistor. With a resistor, current increases linearly as the voltage across it increases. Here, with a
transistor, current from emitter to collector stays limited at a fixed, maximum value no matter how
high the voltage across emitter and collector increases.
Often it is useful to superimpose several collector current/voltage graphs for different base currents on the same graph. A collection of curves like this one curve plotted for each distinct level
of base current for a particular transistor is called the transistors characteristic curves:
9

Ibase = 75 A

8
7

Icollector
(mA)

6
5

Ibase = 40 A

4
3

Ibase = 20 A

Ibase = 5 A

1
0
0

Ecollector-to-emitter (V)

10

11

12

13

14

4.4.

ACTIVE MODE OPERATION

183

Each curve on the graph reflects the collector current of the transistor, plotted over a range of
collector-to-emitter voltages, for a given amount of base current. Since a transistor tends to act as
a current regulator, limiting collector current to a proportion set by the base current, it is useful
to express this proportion as a standard transistor performance measure. Specifically, the ratio of
collector current to base current is known as the Beta ratio (symbolized by the Greek letter ):

Icollector
Ibase

is also known as hfe

Sometimes the ratio is designated as hf e , a label used in a branch of mathematical semiconductor analysis known as hybrid parameters which strives to achieve very precise predictions of
transistor performance with detailed equations. Hybrid parameter variables are many, but they are
all labeled with the general letter h and a specific subscript. The variable h f e is just another
(standardized) way of expressing the ratio of collector current to base current, and is interchangeable
with . Like all ratios, is unitless.

for any transistor is determined by its design: it cannot be altered after manufacture. However,
there are so many physical variables impacting that it is rare to have two transistors of the same
design exactly match. If a circuit design relies on equal ratios between multiple transistors,
matched sets of transistors may be purchased at extra cost. However, it is generally considered
bad design practice to engineer circuits with such dependencies.

It would be nice if the of a transistor remained stable for all operating conditions, but this
is not true in real life. For an actual transistor, the ratio may vary by a factor of over 3 within
its operating current limits. For example, a transistor with advertised of 50 may actually test
with Ic /Ib ratios as low as 30 and as high as 100, depending on the amount of collector current,
the transistors temperature, and frequency of amplified signal, among other factors. For tutorial
purposes it is adequate to assume a constant for any given transistor (which is what SPICE tends
to do in a simulation), but just realize that real life is not that simple!

Sometimes it is helpful for comprehension to model complex electronic components with a


collection of simpler, better-understood components. The following is a popular model shown in
many introductory electronics texts:

184

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

C
B
E
NPN diode-rheostat model
C

This model casts the transistor as a combination of diode and rheostat (variable resistor). Current
through the base-emitter diode controls the resistance of the collector-emitter rheostat (as implied
by the dashed line connecting the two components), thus controlling collector current. An NPN
transistor is modeled in the figure shown, but a PNP transistor would be only slightly different
(only the base-emitter diode would be reversed). This model succeeds in illustrating the basic
concept of transistor amplification: how the base current signal can exert control over the collector
current. However, I personally dont like this model because it tends to miscommunicate the notion
of a set amount of collector-emitter resistance for a given amount of base current. If this were true,
the transistor wouldnt regulate collector current at all like the characteristic curves show. Instead
of the collector current curves flattening out after their brief rise as the collector-emitter voltage
increases, the collector current would be directly proportional to collector-emitter voltage, rising
steadily in a straight line on the graph.

A better transistor model, often seen in more advanced textbooks, is this:

4.4.

ACTIVE MODE OPERATION

185

C
B
E
NPN diode-current source model
C

It casts the transistor as a combination of diode and current source, the output of the current
source being set at a multiple ( ratio) of the base current. This model is far more accurate in
depicting the true input/output characteristics of a transistor: base current establishes a certain
amount of collector current, rather than a certain amount of collector-emitter resistance as the first
model implies. Also, this model is favored when performing network analysis on transistor circuits,
the current source being a well-understood theoretical component. Unfortunately, using a current
source to model the transistors current-controlling behavior can be misleading: in no way will the
transistor ever act as a source of electrical energy, which the current source symbol implies is a
possibility.

My own personal suggestion for a transistor model substitutes a constant-current diode for the
current source:

186

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

C
B
E
NPN diode-regulating diode model
C

E
Since no diode ever acts as a source of electrical energy, this analogy escapes the false implication
of the current source model as a source of power, while depicting the transistors constant-current
behavior better than the rheostat model. Another way to describe the constant-current diodes
action would be to refer to it as a current regulator, so this transistor illustration of mine might also
be described as a diode-current regulator model. The greatest disadvantage I see to this model is the
relative obscurity of constant-current diodes. Many people may be unfamiliar with their symbology
or even of their existence, unlike either rheostats or current sources, which are commonly known.
REVIEW:
A transistor is said to be in its active mode if it is operating somewhere between fully on
(saturated) and fully off (cutoff).
Base current tends to regulate collector current. By regulate, we mean that no more collector
current may exist than what is allowed by the base current.
The ratio between collector current and base current is called Beta () or h f e .
ratios are different for every transistor, and they tend to change for different operating
conditions.

4.5

The common-emitter amplifier

At the beginning of this chapter we saw how transistors could be used as switches, operating in
either their saturation or cutoff modes. In the last section we saw how transistors behave

4.5.

THE COMMON-EMITTER AMPLIFIER

187

within their active modes, between the far limits of saturation and cutoff. Because transistors
are able to control current in an analog (infinitely divisible) fashion, they find use as amplifiers for
analog signals.
One of the simpler transistor amplifier circuits to study is the one used previously for illustrating
the transistors switching ability:

solar
cell
It is called the common-emitter configuration because (ignoring the power supply battery) both
the signal source and the load share the emitter lead as a common connection point. This is not the
only way in which a transistor may be used as an amplifier, as we will see in later sections of this
chapter:

Vout
C

B
solar
cell

Load

Vin

Before, this circuit was shown to illustrate how a relatively small current from a solar cell could be
used to saturate a transistor, resulting in the illumination of a lamp. Knowing now that transistors
are able to throttle their collector currents according to the amount of base current supplied by
an input signal source, we should be able to see that the brightness of the lamp in this circuit is
controllable by the solar cells light exposure. When there is just a little light shone on the solar
cell, the lamp will glow dimly. The lamps brightness will steadily increase as more light falls on the
solar cell.
Suppose that we were interested in using the solar cell as a light intensity instrument. We want
to be able to measure the intensity of incident light with the solar cell by using its output current to
drive a meter movement. It is possible to directly connect a meter movement to a solar cell for this
purpose. In fact, the simplest light-exposure meters for photography work are designed like this:

meter movement

+
solar
cell

188

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

While this approach might work for moderate light intensity measurements, it would not work as
well for low light intensity measurements. Because the solar cell has to supply the meter movements
power needs, the system is necessarily limited in its sensitivity. Supposing that our need here is to
measure very low-level light intensities, we are pressed to find another solution.
Perhaps the most direct solution to this measurement problem is to use a transistor to amplify
the solar cells current so that more meter movement needle deflection may be obtained for less
incident light. Consider this approach:

+
+

+
solar cell

Current through the meter movement in this circuit will be times the solar cell current. With
a transistor of 100, this represents a substantial increase in measurement sensitivity. It is prudent
to point out that the additional power to move the meter needle comes from the battery on the far
right of the circuit, not the solar cell itself. All the solar cells current does is control battery current
to the meter to provide a greater meter reading than the solar cell could provide unaided.
Because the transistor is a current-regulating device, and because meter movement indications
are based on the amount of current through their movement coils, meter indication in this circuit
should depend only on the amount of current from the solar cell, not on the amount of voltage
provided by the battery. This means the accuracy of the circuit will be independent of battery
condition, a significant feature! All that is required of the battery is a certain minimum voltage and
current output ability to be able to drive the meter full-scale if needed.
Another way in which the common-emitter configuration may be used is to produce an output
voltage derived from the input signal, rather than a specific output current. Lets replace the meter
movement with a plain resistor and measure voltage between collector and emitter:

R
+
+
solar cell

Voutput

With the solar cell darkened (no current), the transistor will be in cutoff mode and behave as
an open switch between collector and emitter. This will produce maximum voltage drop between
collector and emitter for maximum Voutput , equal to the full voltage of the battery.
At full power (maximum light exposure), the solar cell will drive the transistor into saturation
mode, making it behave like a closed switch between collector and emitter. The result will be
minimum voltage drop between collector and emitter, or almost zero output voltage. In actuality,
a saturated transistor can never achieve zero voltage drop between collector and emitter due to

4.5.

THE COMMON-EMITTER AMPLIFIER

189

the two PN junctions through which collector current must travel. However, this collector-emitter
saturation voltage will be fairly low, around several tenths of a volt, depending on the specific
transistor used.
For light exposure levels somewhere between zero and maximum solar cell output, the transistor
will be in its active mode, and the output voltage will be somewhere between zero and full battery
voltage. An important quality to note here about the common-emitter configuration is that the
output voltage is inversely proportional to the input signal strength. That is, the output voltage
decreases as the input signal increases. For this reason, the common-emitter amplifier configuration
is referred to as an inverting amplifier.
A quick SPICE simulation will verify our qualitative conclusions about this amplifier circuit:

5 k
1

Q1

I1
0

common-emitter amplifier
i1 0 1 dc
q1 2 1 0 mod1
r 3 2 5000
v1 3 0 dc 15
.model mod1 npn
.dc i1 0 50u 2u
.plot dc v(2,0)
.end

type
is
bf
nf
br
nr

npn
1.00E-16
100.000
1.000
1.000
1.000

Voutput

+
V1

15 V
0

190

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

At the beginning of the simulation where the current source (solar cell) is outputting zero current,
the transistor is in cutoff mode and the full 15 volts from the battery is shown at the amplifier
output (between nodes 2 and 0). As the solar cells current begins to increase, the output voltage
proportionally decreases, until the transistor reaches saturation at 30 A of base current (3 mA
of collector current). Notice how the output voltage trace on the graph is perfectly linear (1 volt
steps from 15 volts to 1 volt) until the point of saturation, where it never quite reaches zero. This
is the effect mentioned earlier, where a saturated transistor can never achieve exactly zero voltage
drop between collector and emitter due to internal junction effects. What we do see is a sharp
output voltage decrease from 1 volt to 0.2261 volts as the input current increases from 28 A to 30
A, and then a continuing decrease in output voltage from then on (albeit in progressively smaller
steps). The lowest the output voltage ever gets in this simulation is 0.1299 volts, asymptotically
approaching zero.
So far, weve seen the transistor used as an amplifier for DC signals. In the solar cell light
meter example, we were interested in amplifying the DC output of the solar cell to drive a DC
meter movement, or to produce a DC output voltage. However, this is not the only way in which
a transistor may be employed as an amplifier. In many cases, what is desired is an AC amplifier
for amplifying alternating current and voltage signals. One common application of this is in audio
electronics (radios, televisions, and public-address systems). Earlier, we saw an example where the
audio output of a tuning fork could be used to activate a transistor as a switch. Lets see if we can
modify that circuit to send power to a speaker rather than to a lamp:

4.5.

THE COMMON-EMITTER AMPLIFIER

191

microphone

source of
sound

In the original circuit, a full-wave bridge rectifier was used to convert the microphones AC
output signal into a DC voltage to drive the input of the transistor. All we cared about here was
turning the lamp on with a sound signal from the microphone, and this arrangement sufficed for
that purpose. But now we want to actually reproduce the AC signal and drive a speaker. This
means we cannot rectify the microphones output anymore, because we need undistorted AC signal
to drive the transistor! Lets remove the bridge rectifier and replace the lamp with a speaker:
speaker

microphone

source of
sound

Since the microphone may produce voltages exceeding the forward voltage drop of the baseemitter PN (diode) junction, Ive placed a resistor in series with the microphone. Lets simulate this
circuit now in SPICE and see what happens:

speaker
8

3
1
Vinput
1.5 V
2 kHz

R1

Q1

V1

15 V

1 k
0

common-emitter amplifier
vinput 1 0 sin (0 1.5 2000 0 0)
r1 1 2 1k

192

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

q1 3 2 0 mod1
rspkr 3 4 8
v1 4 0 dc 15
.model mod1 npn
.tran 0.02m 0.74m
.plot tran v(1,0) i(v1)
.end

The simulation plots both the input voltage (an AC signal of 1.5 volt peak amplitude and 2000
Hz frequency) and the current through the 15 volt battery, which is the same as the current through
the speaker. What we see here is a full AC sine wave alternating in both positive and negative
directions, and a half-wave output current waveform that only pulses in one direction. If we were
actually driving a speaker with this waveform, the sound produced would be horribly distorted.
Whats wrong with the circuit? Why wont it faithfully reproduce the entire AC waveform
from the microphone? The answer to this question is found by close inspection of the transistor
diode-regulating diode model:

4.5.

THE COMMON-EMITTER AMPLIFIER

193

C
B
E
NPN diode-regulating diode model
C

Collector current is controlled, or regulated, through the constant-current mechanism according


to the pace set by the current through the base-emitter diode. Note that both current paths through
the transistor are monodirectional: one way only! Despite our intent to use the transistor to amplify
an AC signal, it is essentially a DC device, capable of handling currents in a single direction only. We
may apply an AC voltage input signal between the base and emitter, but electrons cannot flow in that
circuit during the part of the cycle that reverse-biases the base-emitter diode junction. Therefore,
the transistor will remain in cutoff mode throughout that portion of the cycle. It will turn on in its
active mode only when the input voltage is of the correct polarity to forward-bias the base-emitter
diode, and only when that voltage is sufficiently high to overcome the diodes forward voltage drop.
Remember that bipolar transistors are current-controlled devices: they regulate collector current
based on the existence of base-to-emitter current, not base-to-emitter voltage.

The only way we can get the transistor to reproduce the entire waveform as current through the
speaker is to keep the transistor in its active mode the entire time. This means we must maintain
current through the base during the entire input waveform cycle. Consequently, the base-emitter
diode junction must be kept forward-biased at all times. Fortunately, this can be accomplished with
the aid of a DC bias voltage added to the input signal. By connecting a sufficient DC voltage in
series with the AC signal source, forward-bias can be maintained at all points throughout the wave
cycle:

194

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

speaker
8

3
1
Vinput
1.5 V
2 kHz

R1

Q1

V1

15 V

1 k
+
5
Vbias
2.3 V

common-emitter amplifier
vinput 1 5 sin (0 1.5 2000 0 0)
vbias 5 0 dc 2.3
r1 1 2 1k
q1 3 2 0 mod1
rspkr 3 4 8
v1 4 0 dc 15
.model mod1 npn
.tran 0.02m 0.78m
.plot tran v(1,0) i(v1)
.end

With the bias voltage source of 2.3 volts in place, the transistor remains in its active mode
throughout the entire cycle of the wave, faithfully reproducing the waveform at the speaker. Notice

4.5.

THE COMMON-EMITTER AMPLIFIER

195

that the input voltage (measured between nodes 1 and 0) fluctuates between about 0.8 volts and
3.8 volts, a peak-to-peak voltage of 3 volts just as expected (source voltage = 1.5 volts peak). The
output (speaker) current varies between zero and almost 300 mA, 180o out of phase with the input
(microphone) signal.
The following illustration is another view of the same circuit, this time with a few oscilloscopes
(scopemeters) connected at crucial points to display all the pertinent signals:

speaker
8
R1
Vinput
1.5 V
2 kHz

Q1

V1

15 V

1 k
+

Vbias

The need for biasing a transistor amplifier circuit to obtain full waveform reproduction is an
important consideration. A separate section of this chapter will be devoted entirely to the subject
biasing and biasing techniques. For now, it is enough to understand that biasing may be necessary
for proper voltage and current output from the amplifier.
Now that we have a functioning amplifier circuit, we can investigate its voltage, current, and
power gains. The generic transistor used in these SPICE analyses has a of 100, as indicated by
the short transistor statistics printout included in the text output (these statistics were cut from the
last two analyses for brevitys sake):
type
is
bf
nf
br
nr

npn
1.00E-16
100.000
1.000
1.000
1.000

is listed under the abbreviation bf, which actually stands for beta, f orward. If we wanted
to insert our own ratio for an analysis, we could have done so on the .model line of the SPICE
netlist.
Since is the ratio of collector current to base current, and we have our load connected in series
with the collector terminal of the transistor and our source connected in series with the base, the
ratio of output current to input current is equal to beta. Thus, our current gain for this example
amplifier is 100, or 40 dB.

196

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

Voltage gain is a little more complicated to figure than current gain for this circuit. As always,
voltage gain is defined as the ratio of output voltage divided by input voltage. In order to experimentally determine this, we need to modify our last SPICE analysis to plot output voltage rather
than output current so we have two voltage plots to compare:
common-emitter amplifier
vinput 1 5 sin (0 1.5 2000 0 0)
vbias 5 0 dc 2.3
r1 1 2 1k
q1 3 2 0 mod1
rspkr 3 4 8
v1 4 0 dc 15
.model mod1 npn
.tran 0.02m 0.78m
.plot tran v(1,0) v(4,3)
.end

Plotted on the same scale (from 0 to 4 volts), we see that the output waveform (+) has a
smaller peak-to-peak amplitude than the input waveform (*), in addition to being at a lower bias
voltage, not elevated up from 0 volts like the input. Since voltage gain for an AC amplifier is defined
by the ratio of AC amplitudes, we can ignore any DC bias separating the two waveforms. Even so,
the input waveform is still larger than the output, which tells us that the voltage gain is less than 1
(a negative dB figure).
To be honest, this low voltage gain is not characteristic to all common-emitter amplifiers. In this
case it is a consequence of the great disparity between the input and load resistances. Our input
resistance (R1 ) here is 1000 , while the load (speaker) is only 8 . Because the current gain of
this amplifier is determined solely by the of the transistor, and because that figure is fixed, the
current gain for this amplifier wont change with variations in either of these resistances. However,

4.5.

THE COMMON-EMITTER AMPLIFIER

197

voltage gain is dependent on these resistances. If we alter the load resistance, making it a larger
value, it will drop a proportionately greater voltage for its range of load currents, resulting in a
larger output waveform. Lets try another simulation, only this time with a 30 load instead of an
8 load:
common-emitter amplifier
vinput 1 5 sin (0 1.5 2000 0 0)
vbias 5 0 dc 2.3
r1 1 2 1k
q1 3 2 0 mod1
rspkr 3 4 30
v1 4 0 dc 15
.model mod1 npn
.tran 0.02m 0.78m
.plot tran v(1,0) v(4,3)
.end

This time the output voltage waveform is significantly greater in amplitude than the input
waveform. Looking closely, we can see that the output waveform (+) crests between 0 and about
9 volts: approximately 3 times the amplitude of the input voltage.
We can perform another computer analysis of this circuit, only this time instructing SPICE to
analyze it from an AC point of view, giving us peak voltage figures for input and output instead of
a time-based plot of the waveforms:
common-emitter amplifier
vinput 1 5 ac 1.5
vbias 5 0 dc 2.3

198

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

r1 1 2 1k
q1 3 2 0 mod1
rspkr 3 4 30
v1 4 0 dc 15
.model mod1 npn
.ac lin 1 2000 2000
.print ac v(1,0) v(4,3)
.end
freq
2.000E+03

v(1)
1.500E+00

v(4,3)
4.418E+00

Peak voltage measurements of input and output show an input of 1.5 volts and an output of
4.418 volts. This gives us a voltage gain ratio of 2.9453 (4.418 V / 1.5 V), or 9.3827 dB.

AV =

Vout
Vin

AV = 4.418 V
1.5 V
AV = 2.9453
AV(dB) = 20 log AV(ratio)
AV(dB) = 20 log 2.9453
AV(dB) = 9.3827 dB
Because the current gain of the common-emitter amplifier is fixed by , and since the input
and output voltages will be equal to the input and output currents multiplied by their respective
resistors, we can derive an equation for approximate voltage gain:

AV =

Rout
Rin

AV = (100)

30
1000

AV = 3
AV(dB) = 20 log AV(ratio)
AV(dB) = 20 log 3
AV(dB) = 9.5424 dB

4.5.

THE COMMON-EMITTER AMPLIFIER

199

As you can see, the predicted results for voltage gain are quite close to the simulated results.
With perfectly linear transistor behavior, the two sets of figures would exactly match. SPICE does
a reasonable job of accounting for the many quirks of bipolar transistor function in its analysis,
hence the slight mismatch in voltage gain based on SPICEs output.
These voltage gains remain the same regardless of where we measure output voltage in the
circuit: across collector and emitter, or across the series load resistor as we did in the last analysis.
The amount of output voltage change for any given amount of input voltage will remain the same.
Consider the two following SPICE analyses as proof of this. The first simulation is time-based, to
provide a plot of input and output voltages. You will notice that the two signals are 180 o out of
phase with each other. The second simulation is an AC analysis, to provide simple, peak voltage
readings for input and output:
common-emitter amplifier
vinput 1 5 sin (0 1.5 2000 0 0)
vbias 5 0 dc 2.3
r1 1 2 1k
q1 3 2 0 mod1
rspkr 3 4 30
v1 4 0 dc 15
.model mod1 npn
.tran 0.02m 0.74m
.plot tran v(1,0) v(3,0)
.end

common-emitter amplifier
vinput 1 5 ac 1.5
vbias 5 0 dc 2.3

200

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

r1 1 2 1k
q1 3 2 0 mod1
rspkr 3 4 30
v1 4 0 dc 15
.model mod1 npn
.ac lin 1 2000 2000
.print ac v(1,0) v(3,0)
.end
freq
2.000E+03

v(1)
1.500E+00

v(3)
4.418E+00

We still have a peak output voltage of 4.418 volts with a peak input voltage of 1.5 volts. The
only difference from the last set of simulations is the phase of the output voltage.
So far, the example circuits shown in this section have all used NPN transistors. PNP transistors
are just as valid to use as NPN in any amplifier configuration, so long as the proper polarity and
current directions are maintained, and the common-emitter amplifier is no exception. The inverting
behavior and gain properties of a PNP transistor amplifier are the same as its NPN counterpart,
just the polarities are different:

Vinput
-

+
Vbias

REVIEW:
Common-emitter transistor amplifiers are so-called because the input and output voltage
points share the emitter lead of the transistor in common with each other, not considering
any power supplies.
Transistors are essentially DC devices: they cannot directly handle voltages or currents that
reverse direction. In order to make them work for amplifying AC signals, the input signal must
be offset with a DC voltage to keep the transistor in its active mode throughout the entire
cycle of the wave. This is called biasing.

4.6.

THE COMMON-COLLECTOR AMPLIFIER

201

If the output voltage is measured between emitter and collector on a common-emitter amplifier,
it will be 180o out of phase with the input voltage waveform. For this reason, the commonemitter amplifier is called an inverting amplifier circuit.
The current gain of a common-emitter transistor amplifier with the load connected in series
with the collector is equal to . The voltage gain of a common-emitter transistor amplifier is
approximately given here:

AV =

Rout
Rin

Where Rout is the resistor connected in series with the collector and Rin is the resistor
connected in series with the base.

4.6

The common-collector amplifier

Our next transistor configuration to study is a bit simpler in terms of gain calculations. Called the
common-collector configuration, its schematic diagram looks like this:

Vin

Vout

Rload

It is called the common-collector configuration because (ignoring the power supply battery) both
the signal source and the load share the collector lead as a common connection point:

C
B
E
Vin

Vout

Rload

It should be apparent that the load resistor in the common-collector amplifier circuit receives
both the base and collector currents, being placed in series with the emitter. Since the emitter
lead of a transistor is the one handling the most current (the sum of base and collector currents,

202

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

since base and collector currents always mesh together to form the emitter current), it would be
reasonable to presume that this amplifier will have a very large current gain (maximum output
current for minimum input current). This presumption is indeed correct: the current gain for a
common-collector amplifier is quite large, larger than any other transistor amplifier configuration.
However, this is not necessarily what sets it apart from other amplifier designs.
Lets proceed immediately to a SPICE analysis of this amplifier circuit, and you will be able to
immediately see what is unique about this amplifier:

2
1

V1

Vin

15 V

Rload
5 k
0

common-collector amplifier
vin 1 0
q1 2 1 3 mod1
v1 2 0 dc 15
rload 3 0 5k
.model mod1 npn
.dc vin 0 5 0.2
.plot dc v(3,0)
.end

type
is
bf
nf
br
nr

Q1

npn
1.00E-16
100.000
1.000
1.000
1.000

4.6.

THE COMMON-COLLECTOR AMPLIFIER

203

Unlike the common-emitter amplifier from the previous section, the common-collector produces
an output voltage in direct rather than inverse proportion to the rising input voltage. As the input
voltage increases, so does the output voltage. More than that, a close examination reveals that the
output voltage is nearly identical to the input voltage, lagging behind only about 0.77 volts.
This is the unique quality of the common-collector amplifier: an output voltage that is nearly
equal to the input voltage. Examined from the perspective of output voltage change for a given
amount of input voltage change, this amplifier has a voltage gain of almost exactly unity (1), or 0
dB. This holds true for transistors of any value, and for load resistors of any resistance value.
It is simple to understand why the output voltage of a common-collector amplifier is always
nearly equal to the input voltage. Referring back to the diode-regulating diode transistor model,
we see that the base current must go through the base-emitter PN junction, which is equivalent
to a normal rectifying diode. So long as this junction is forward-biased (the transistor conducting
current in either its active or saturated modes), it will have a voltage drop of approximately 0.7 volts,
assuming silicon construction. This 0.7 volt drop is largely irrespective of the actual magnitude of
base current, so we can regard it as being constant:

204

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

C
B

+
0.7 V
-

Vin

E
+
Rload
-

Given the voltage polarities across the base-emitter PN junction and the load resistor, we see
that they must add together to equal the input voltage, in accordance with Kirchhoffs Voltage Law.
In other words, the load voltage will always be about 0.7 volts less than the input voltage for all
conditions where the transistor is conducting. Cutoff occurs at input voltages below 0.7 volts, and
saturation at input voltages in excess of battery (supply) voltage plus 0.7 volts.
Because of this behavior, the common-collector amplifier circuit is also known as the voltagefollower or emitter-follower amplifier, in reference to the fact that the input and load voltages follow
each other so closely.
Applying the common-collector circuit to the amplification of AC signals requires the same input
biasing used in the common-emitter circuit: a DC voltage must be added to the AC input signal
to keep the transistor in its active mode during the entire cycle. When this is done, the result is a
non-inverting amplifier:

2
1

V1

1.5 V
2 kHz

Vin

Q1

15 V

Rload
5 k
-

4
Vbias
2.3 V

common-collector amplifier
vin 1 4 sin(0 1.5 2000 0 0)
vbias 4 0 dc 2.3
q1 2 1 3 mod1
v1 2 0 dc 15
rload 3 0 5k

4.6.

THE COMMON-COLLECTOR AMPLIFIER

205

.model mod1 npn


.tran .02m .78m
.plot tran v(1,0) v(3,0)
.end

Heres another view of the circuit, this time with oscilloscopes connected to several points of
interest:

V1
Vin

1.5 V
2 kHz
+

15 V

Rload
5 k
-

Since this amplifier configuration doesnt provide any voltage gain (in fact, in practice it actually
has a voltage gain of slightly less than 1), its only amplifying factor is current. The common-emitter
amplifier configuration examined in the previous section had a current gain equal to the of the
transistor, being that the input current went through the base and the output (load) current went

206

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

through the collector, and by definition is the ratio between the collector and base currents. In
the common-collector configuration, though, the load is situated in series with the emitter, and thus
its current is equal to the emitter current. With the emitter carrying collector current and base
current, the load in this type of amplifier has all the current of the collector running through it plus
the input current of the base. This yields a current gain of plus 1:

AI =

Iemitter
Ibase

AI =

Icollector+ Ibase
Ibase

AI =

Icollector
+1
Ibase

AI = + 1
Once again, PNP transistors are just as valid to use in the common-collector configuration as
NPN transistors. The gain calculations are all the same, as is the non-inverting behavior of the
amplifier. The only difference is in voltage polarities and current directions:

Vin

Rload
-

A popular application of the common-collector amplifier is for regulated DC power supplies,


where an unregulated (varying) source of DC voltage is clipped at a specified level to supply regulated
(steady) voltage to a load. Of course, zener diodes already provide this function of voltage regulation:

R
Unregulated
DC voltage
source

Zener
diode

Rload

Regulated voltage
across load

4.6.

THE COMMON-COLLECTOR AMPLIFIER

207

However, when used in this direct fashion, the amount of current that may be supplied to the
load is usually quite limited. In essence, this circuit regulates voltage across the load by keeping
current through the series resistor at a high enough level to drop all the excess power source voltage
across it, the zener diode drawing more or less current as necessary to keep the voltage across itself
steady. For high-current loads, an plain zener diode voltage regulator would have to be capable of
shunting a lot of current through the diode in order to be effective at regulating load voltage in the
event of large load resistance or voltage source changes.
One popular way to increase the current-handling ability of a regulator circuit like this is to use
a common-collector transistor to amplify current to the load, so that the zener diode circuit only
has to handle the amount of current necessary to drive the base of the transistor:

R
Unregulated
DC voltage
source

Zener
diode

Rload

Theres really only one caveat to this approach: the load voltage will be approximately 0.7 volts
less than the zener diode voltage, due to the transistors 0.7 volt base-emitter drop. However, since
this 0.7 volt difference is fairly constant over a wide range of load currents, a zener diode with a 0.7
volt higher rating can be chosen for the application.
Sometimes the high current gain of a single-transistor, common-collector configuration isnt
enough for a particular application. If this is the case, multiple transistors may be staged together
in a popular configuration known as a Darlington pair, just an extension of the common-collector
concept:

An NPN "Darlington pair"


C

208

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

Darlington pairs essentially place one transistor as the common-collector load for another transistor, thus multiplying their individual current gains. Base current through the upper-left transistor is
amplified through that transistors emitter, which is directly connected to the base of the lower-right
transistor, where the current is again amplified. The overall current gain is as follows:

Darlington pair current gain


AI = (1 + 1)(2 + 1)
Where,
1 = Beta of first transistor
2 = Beta of second transistor
Voltage gain is still nearly equal to 1 if the entire assembly is connected to a load in commoncollector fashion, although the load voltage will be a full 1.4 volts less than the input voltage:

+
0.7 V +

+
0.7 V -

Rload

Vin

Vout

Vout = Vin - 1.4


Darlington pairs may be purchased as discrete units (two transistors in the same package), or may
be built up from a pair of individual transistors. Of course, if even more current gain is desired than
what may be obtained with a pair, Darlington triplet or quadruplet assemblies may be constructed.
REVIEW:
Common-collector transistor amplifiers are so-called because the input and output voltage
points share the collector lead of the transistor in common with each other, not considering
any power supplies.
The output voltage on a common-collector amplifier will be in phase with the input voltage,
making the common-collector a non-inverting amplifier circuit.
The current gain of a common-collector amplifier is equal to plus 1. The voltage gain is
approximately equal to 1 (in practice, just a little bit less).
A Darlington pair is a pair of transistors piggybacked on one another so that the emitter of
one feeds current to the base of the other in common-collector form. The result is an overall

4.7.

THE COMMON-BASE AMPLIFIER

209

current gain equal to the product (multiplication) of their individual common-collector current
gains ( plus 1).

4.7

The common-base amplifier

The final transistor amplifier configuration we need to study is the common-base. This configuration
is more complex than the other two, and is less common due to its strange operating characteristics.

Rload

Vin

It is called the common-base configuration because (DC power source aside), the signal source
and the load share the base of the transistor as a common connection point:

C
B

Vin

Rload V
out

Perhaps the most striking characteristic of this configuration is that the input signal source
must carry the full emitter current of the transistor, as indicated by the heavy arrows in the first
illustration. As we know, the emitter current is greater than any other current in the transistor,
being the sum of base and collector currents. In the last two amplifier configurations, the signal
source was connected to the base lead of the transistor, thus handling the least current possible.
Because the input current exceeds all other currents in the circuit, including the output current,
the current gain of this amplifier is actually less than 1 (notice how Rload is connected to the collector,
thus carrying slightly less current than the signal source). In other words, it attenuates current
rather than amplifying it. With common-emitter and common-collector amplifier configurations,
the transistor parameter most closely associated with gain was . In the common-base circuit, we
follow another basic transistor parameter: the ratio between collector current and emitter current,
which is a fraction always less than 1. This fractional value for any transistor is called the alpha
ratio, or ratio.
Since it obviously cant boost signal current, it only seems reasonable to expect it to boost signal
voltage. A SPICE simulation will vindicate that assumption:

210

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

2
R1
100
Vin
+
1

Q1

Rload

V1

5.0 k
+ 3

15 V
common-base amplifier
vin 0 1
r1 1 2 100
q1 4 0 2 mod1
v1 3 0 dc 15
rload 3 4 5k
.model mod1 npn
.dc vin 0.6 1.2 .02
.plot dc v(3,4)
.end

Notice how in this simulation the output voltage goes from practically nothing (cutoff) to 15.75
volts (saturation) with the input voltage being swept over a range of 0.6 volts to 1.2 volts. In fact,
the output voltage plot doesnt show a rise until about 0.7 volts at the input, and cuts off (flattens)
at about 1.12 volts input. This represents a rather large voltage gain with an output voltage span
of 15.75 volts and an input voltage span of only 0.42 volts: a gain ratio of 37.5, or 31.48 dB. Notice
also how the output voltage (measured across Rload ) actually exceeds the power supply (15 volts)
at saturation, due to the series-aiding effect of the the input voltage source.

4.7.

THE COMMON-BASE AMPLIFIER

211

A second set of SPICE analyses with an AC signal source (and DC bias voltage) tells the same
story: a high voltage gain.

Vin
2
100

Q1
Rload
5.0 k

0.12 Vp-p
0 Voffset
2 kHz

R1

Vbias 0.95 V

V1

+ 3

15 V

common-base amplifier
vin 5 2 sin (0 0.12 2000 0 0)
vbias 0 1 dc 0.95
r1 2 1 100
q1 4 0 5 mod1
v1 3 0 dc 15
rload 3 4 5k
.model mod1 npn
.tran 0.02m 0.78m
.plot tran v(5,2) v(4,3)
.end

Vout

212

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

As you can see, the input and output waveforms are in phase with each other. This tells us that
the common-base amplifier is non-inverting.

common-base amplifier
vin 5 2 sin (0 0.12 2000 0 0)
vbias 0 1 dc 0.95
r1 2 1 100
q1 4 0 5 mod1
v1 3 0 dc 15
rload 3 4 5k
.model mod1 npn
.ac lin 1 2000 2000
.print ac v(5,2) v(3,4)
.end

freq
2.000E+03

v(1)
1.200E-01

v(3,4)
5.129E+00

Voltage figures from the second analysis (AC mode) show a voltage gain of 42.742 (5.129 V /
0.12 V), or 32.617 dB:

AV =

Vout
Vin

AV = 5.129 V
0.12 V
AV = 42.742
AV(dB) = 20 log AV(ratio)
AV(dB) = 20 log 42.742
AV(dB) = 32.617 dB
Heres another view of the circuit, showing the phase relations and DC offsets of various signals
in the circuit just simulated:

4.7.

THE COMMON-BASE AMPLIFIER

213

Vin
Q1
Rload

V1

Vbias
. . . and for a PNP transistor:

Vin
Q1
Rload

Vbias
+

V1

Predicting voltage gain for the common-base amplifier configuration is quite difficult, and involves approximations of transistor behavior that are difficult to measure directly. Unlike the other
amplifier configurations, where voltage gain was either set by the ratio of two resistors (commonemitter), or fixed at an unchangeable value (common-collector), the voltage gain of the common-base
amplifier depends largely on the amount of DC bias on the input signal. As it turns out, the internal
transistor resistance between emitter and base plays a major role in determining voltage gain, and
this resistance changes with different levels of current through the emitter.
While this phenomenon is difficult to explain, it is rather easy to demonstrate through the use of
computer simulation. What Im going to do here is run several SPICE simulations on a common-base
amplifier circuit, changing the DC bias voltage slightly while keeping the AC signal amplitude and
all other circuit parameters constant. As the voltage gain changes from one simulation to another,
different output voltage amplitudes will be noticed as a result.
Although these analyses will all be conducted in the AC mode, they were first proofed in
the transient analysis mode (voltage plotted over time) to ensure that the entire wave was being
faithfully reproduced and not clipped due to improper biasing. No meaningful calculations of gain
can be based on waveforms that are distorted:
common-base amplifier
vin 5 2 sin (0 0.12 2000 0 0)
vbias 0 1 dc 0.95
r1 2 1 100
q1 4 0 5 mod1
v1 3 0 dc 15

214

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

rload 3 4 5k
.model mod1 npn
.ac lin 1 2000 2000
.print ac v(5,2) v(3,4)
.end
freq
2.000E+03

v(5,2)
8.000E-02

v(3,4)
3.005E+00

common-base amplifier
vin 5 2 sin (0 0.12 2000 0 0)
vbias 0 1 dc 0.95
r1 2 1 100
q1 4 0 5 mod1
v1 3 0 dc 15
rload 3 4 5k
.model mod1 npn
.ac lin 1 2000 2000
.print ac v(5,2) v(3,4)
.end
freq
2.000E+03

v(5,2)
8.000E-02

v(3,4)
3.264E+00

common-base amplifier
vin 5 2 sin (0 0.12 2000 0 0)
vbias 0 1 dc 0.95
r1 2 1 100
q1 4 0 5 mod1
v1 3 0 dc 15
rload 3 4 5k
.model mod1 npn
.ac lin 1 2000 2000
.print ac v(5,2) v(3,4)
.end
freq
2.000E+03

v(5,2)
8.000E-02

v(3,4)
3.419E+00

A trend should be evident here: with increases in DC bias voltage, voltage gain increases as well.
We can see that the voltage gain is increasing because each subsequent simulation produces greater
output voltage for the exact same input signal voltage (0.08 volts). As you can see, the changes are
quite large, and they are caused by miniscule variations in bias voltage!
The combination of very low current gain (always less than 1) and somewhat unpredictable
voltage gain conspire against the common-base design, relegating it to few practical applications.
REVIEW:

4.8.

BIASING TECHNIQUES

215

Common-base transistor amplifiers are so-called because the input and output voltage points
share the base lead of the transistor in common with each other, not considering any power
supplies.

The current gain of a common-base amplifier is always less than 1. The voltage gain is a
function of input and output resistances, and also the internal resistance of the emitter-base
junction, which is subject to change with variations in DC bias voltage. Suffice to say that the
voltage gain of a common-base amplifier can be very high.

The ratio of a transistors collector current to emitter current is called . The value for any
transistor is always less than unity, or in other words, less than 1.

4.8

Biasing techniques

In the common-emitter section of this chapter, we saw a SPICE analysis where the output waveform
resembled a half-wave rectified shape: only half of the input waveform was reproduced, with the
other half being completely cut off. Since our purpose at that time was to reproduce the entire
waveshape, this constituted a problem. The solution to this problem was to add a small bias voltage
to the amplifier input so that the transistor stayed in active mode throughout the entire wave cycle.
This addition was called a bias voltage.
There are applications, though, where a half-wave output is not problematic. In fact, some
applications may necessitate this very type of amplification. Because it is possible to operate an
amplifier in modes other than full-wave reproduction, and because there are specific applications
requiring different ranges of reproduction, it is useful to describe the degree to which an amplifier
reproduces the input waveform by designating it according to class. Amplifier class operation is
categorized by means of alphabetical letters: A, B, C, and AB.
Class A operation is where the entire input waveform is faithfully reproduced. Although I didnt
introduce this concept back in the common-emitter section, this is what we were hoping to attain in
our simulations. Class A operation can only be obtained when the transistor spends its entire time
in the active mode, never reaching either cutoff or saturation. To achieve this, sufficient DC bias
voltage is usually set at the level necessary to drive the transistor exactly halfway between cutoff
and saturation. This way, the AC input signal will be perfectly centered between the amplifiers
high and low signal limit levels.

216

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

Class A operation

Vinput
Amplifier
Vbias

Class B operation is what we had the first time an AC signal was applied to the common-emitter
amplifier with no DC bias voltage. The transistor spent half its time in active mode and the other
half in cutoff with the input voltage too low (or even of the wrong polarity!) to forward-bias its
base-emitter junction.

Class B operation

Vinput
Amplifier

Little or no
DC bias voltage
By itself, an amplifier operating in class B mode is not very useful. In most circumstances, the
severe distortion introduced into the waveshape by eliminating half of it would be unacceptable.
However, class B operation is a useful mode of biasing if two amplifiers are operated as a push-pull
pair, each amplifier handling only half of the waveform at a time:

4.8.

BIASING TECHNIQUES

217

Q1

Input components
omitted for simplicity

Power
supply

Voutput
Q2

Transistor Q1 pushes (drives the output voltage in a positive direction with respect to ground),
while transistor Q2 pulls the output voltage (in a negative direction, toward 0 volts with respect
to ground). Individually, each of these transistors is operating in class B mode, active only for
one-half of the input waveform cycle. Together, however, they function as a team to produce an
output waveform identical in shape to the input waveform.

A decided advantage of the class B (push-pull) amplifier design over the class A design is greater
output power capability. With a class A design, the transistor dissipates a lot of energy in the form
of heat because it never stops conducting current. At all points in the wave cycle it is in the active
(conducting) mode, conducting substantial current and dropping substantial voltage. This means
there is substantial power dissipated by the transistor throughout the cycle. In a class B design, each
transistor spends half the time in cutoff mode, where it dissipates zero power (zero current = zero
power dissipation). This gives each transistor a time to rest and cool while the other transistor
carries the burden of the load. Class A amplifiers are simpler in design, but tend to be limited to
low-power signal applications for the simple reason of transistor heat dissipation.

There is another class of amplifier operation known as class AB, which is somewhere between
class A and class B: the transistor spends more than 50% but less than 100% of the time conducting
current.

If the input signal bias for an amplifier is slightly negative (opposite of the bias polarity for
class A operation), the output waveform will be further clipped than it was with class B biasing,
resulting in an operation where the transistor spends the majority of the time in cutoff mode:

218

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

Class C operation

Vinput
Amplifier
Vbias

At first, this scheme may seem utterly pointless. After all, how useful could an amplifier be if it
clips the waveform as badly as this? If the output is used directly with no conditioning of any kind,
it would indeed be of questionable utility. However, with the application of a tank circuit (parallel
resonant inductor-capacitor combination) to the output, the occasional output surge produced by
the amplifier can set in motion a higher-frequency oscillation maintained by the tank circuit. This
may be likened to a machine where a heavy flywheel is given an occasional kick to keep it spinning:

Class C operation
with resonant output

Vinput
Amplifier
Vbias

Called class C operation, this scheme also enjoys high power efficiency due to the fact that
the transistor(s) spend the vast majority of time in the cutoff mode, where they dissipate zero
power. The rate of output waveform decay (decreasing oscillation amplitude between kicks from
the amplifier) is exaggerated here for the benefit of illustration. Because of the tuned tank circuit
on the output, this type of circuit is usable only for amplifying signals of definite, fixed frequency.
Another type of amplifier operation, significantly different from Class A, B, AB, or C, is called
Class D. It is not obtained by applying a specific measure of bias voltage as are the other classes of
operation, but requires a radical re-design of the amplifier circuit itself. Its a little too early in this

4.8.

BIASING TECHNIQUES

219

chapter to investigate exactly how a class D amplifier is built, but not too early to discuss its basic
principle of operation.
A class D amplifier reproduces the profile of the input voltage waveform by generating a rapidlypulsing squarewave output. The duty cycle of this output waveform (time on versus total cycle
time) varies with the instantaneous amplitude of the input signal. The following plots demonstrate
this principle:
Class D operation

Input

Output

The greater the instantaneous voltage of the input signal, the greater the duty cycle of the output
squarewave pulse. If there can be any goal stated of the class D design, it is to avoid active-mode
transistor operation. Since the output transistor of a class D amplifier is never in the active mode,
only cutoff or saturated, there will be little heat energy dissipated by it. This results in very high
power efficiency for the amplifier. Of course, the disadvantage of this strategy is the overwhelming
presence of harmonics on the output. Fortunately, since these harmonic frequencies are typically
much greater than the frequency of the input signal, they can be filtered out by a low-pass filter
with relative ease, resulting in an output more closely resembling the original input signal waveform.
Class D technology is typically seen where extremely high power levels and relatively low frequencies
are encountered, such as in industrial inverters (devices converting DC into AC power to run motors
and other large devices) and high-performance audio amplifiers.
A term you will likely come across in your studies of electronics is something called quiescent,
which is a modifier designating the normal, or zero input signal, condition of a circuit. Quiescent
current, for example, is the amount of current in a circuit with zero input signal voltage applied. Bias
voltage in a transistor circuit forces the transistor to operate at a different level of collector current
with zero input signal voltage than it would without that bias voltage. Therefore, the amount of
bias in an amplifier circuit determines its quiescent values.
In a class A amplifier, the quiescent current should be exactly half of its saturation value (halfway
between saturation and cutoff, cutoff by definition being zero). Class B and class C amplifiers have
quiescent current values of zero, since they are supposed to be cutoff with no signal applied. Class
AB amplifiers have very low quiescent current values, just above cutoff. To illustrate this graphically,

220

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

a load line is sometimes plotted over a transistors characteristic curves to illustrate its range of
operation while connected to a load resistance of specific value:

Ibase = 75 A
"Load line"
Icollector

Ibase = 40 A
Ibase = 20 A
Ibase = 5 A
0

Ecollector-to-emitter
Vsupply

A load line is a plot of collector-to-emitter voltage over a range of base currents. At the lowerright corner of the load line, voltage is at maximum and current is at zero, representing a condition
of cutoff. At the upper-left corner of the line, voltage is at zero while current is at a maximum,
representing a condition of saturation. Dots marking where the load line intersects the various
transistor curves represent realistic operating conditions for those base currents given.
Quiescent operating conditions may be shown on this type of graph in the form of a single dot
along the load line. For a class A amplifier, the quiescent point will be in the middle of the load
line, like this:

Ibase = 75 A

Icollector

Ibase = 40 A

Quiescent point
for class A
operation

Ibase = 20 A
Ibase = 5 A
0

Ecollector-to-emitter
Vsupply

4.8.

BIASING TECHNIQUES

221

In this illustration, the quiescent point happens to fall on the curve representing a base current
of 40 A. If we were to change the load resistance in this circuit to a greater value, it would affect
the slope of the load line, since a greater load resistance would limit the maximum collector current
at saturation, but would not change the collector-emitter voltage at cutoff. Graphically, the result
is a load line with a different upper-left point and the same lower-right point:

Ibase = 75 A

Icollector

The nonhorizontal
portion of
the curve
represents
transistor
saturation

Ibase = 40 A
Ibase = 20 A
Ibase = 5 A
0

Ecollector-to-emitter
Vsupply

Note how the new load line doesnt intercept the 75 A curve along its flat portion as before. This
is very important to realize because the non-horizontal portion of a characteristic curve represents
a condition of saturation. Having the load line intercept the 75 A curve outside of the curves
horizontal range means that the amplifier will be saturated at that amount of base current. Increasing
the load resistor value is what caused the load line to intercept the 75 A curve at this new point,
and it indicates that saturation will occur at a lesser value of base current than before.
With the old, lower-value load resistor in the circuit, a base current of 75 A would yield a
proportional collector current (base current multiplied by ). In the first load line graph, a base
current of 75 A gave a collector current almost twice what was obtained at 40 A, as the ratio
would predict. Now, however, there is only a marginal increase in collector current between base
current values of 75 A and 40 A, because the transistor begins to lose sufficient collector-emitter
voltage to continue to regulate collector current.
In order to maintain linear (no-distortion) operation, transistor amplifiers shouldnt be operated
at points where the transistor will saturate; that is, in any case where the load line will not potentially
fall on the horizontal portion of a collector current curve. In this case, wed have to add a few more
curves to the graph before we could tell just how far we could push this transistor with increased
base currents before it saturates.

222

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

Ibase = 75 A
Ibase = 60 A
Ibase = 50 A

Icollector

Ibase = 40 A
Ibase = 20 A
Ibase = 5 A
0

Ecollector-to-emitter
Vsupply

It appears in this graph that the highest-current point on the load line falling on the straight
portion of a curve is the point on the 50 A curve. This new point should be considered the maximum
allowable input signal level for class A operation. Also for class A operation, the bias should be set
so that the quiescent point is halfway between this new maximum point and cutoff:

Ibase = 75 A
Ibase = 60 A
Ibase = 50 A

Icollector

Ibase = 40 A
Ibase = 20 A

New quiescent point

Ibase = 5 A
0

Ecollector-to-emitter
Vsupply

Now that we know a little more about the consequences of different DC bias voltage levels, it
is time to investigate practical biasing techniques. So far, Ive shown a small DC voltage source
(battery) connected in series with the AC input signal to bias the amplifier for whatever desired
class of operation. In real life, the connection of a precisely-calibrated battery to the input of an
amplifier is simply not practical. Even if it were possible to customize a battery to produce just
the right amount of voltage for any given bias requirement, that battery would not remain at its

4.8.

BIASING TECHNIQUES

223

manufactured voltage indefinitely. Once it started to discharge and its output voltage drooped, the
amplifier would begin to drift in the direction of class B operation.
Take this circuit, illustrated in the common-emitter section for a SPICE simulation, for instance:

speaker
8

3
R1

1
Vinput
1.5 V
2 kHz

Q1

V1

15 V

1 k
+

5
Vbias
2.3 V

That 2.3 volt Vbias battery would not be practical to include in a real amplifier circuit. A far
more practical method of obtaining bias voltage for this amplifier would be to develop the necessary
2.3 volts using a voltage divider network connected across the 15 volt battery. After all, the 15 volt
battery is already there by necessity, and voltage divider circuits are very easy to design and build.
Lets see how this might look:

speaker
4
R2

R1 2

1
Vinput
1.5 V
2 kHz

1 k
Vbias

R3
0

Q1

V1

15 V

If we choose a pair of resistor values for R2 and R3 that will produce 2.3 volts across R3 from a
total of 15 volts (such as 8466 for R2 and 1533 for R3 ), we should have our desired value of 2.3
volts between base and emitter for biasing with no signal input. The only problem is, this circuit
configuration places the AC input signal source directly in parallel with R 3 of our voltage divider.
This is not acceptable, as the AC source will tend to overpower any DC voltage dropped across R 3 .
Parallel components must have the same voltage, so if an AC voltage source is directly connected
across one resistor of a DC voltage divider, the AC source will win and there will be no DC bias
voltage added to the signal.
One way to make this scheme work, although it may not be obvious why it will work, is to place
a coupling capacitor between the AC voltage source and the voltage divider like this:

224

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

speaker
4
C
1
Vinput
1.5 V
2 kHz
0

R2

8.466 k
5 R1 2

R3

1 k
1.533 k

8
V1

Q1

15 V

The capacitor forms a high-pass filter between the AC source and the DC voltage divider, passing
almost all of the AC signal voltage on to the transistor while blocking all DC voltage from being
shorted through the AC signal source. This makes much more sense if you understand the superposition theorem and how it works. According to superposition, any linear, bilateral circuit can be
analyzed in a piecemeal fashion by only considering one power source at a time, then algebraically
adding the effects of all power sources to find the final result. If we were to separate the capacitor
and R2 R3 voltage divider circuit from the rest of the amplifier, it might be easier to understand
how this superposition of AC and DC would work.
With only the AC signal source in effect, and a capacitor with an arbitrarily low impedance at
signal frequency, almost all the AC voltage appears across R3 :

1.5 V
2 kHz

Vinput
1.5 V
2 kHz

R2

8.466
k

R3

1.533
k

Due to the capacitors very low impedance at


signal frequency, it behaves much like a straight
piece of wire and thus can be omitted for the
purpose of this step in superposition analysis.
With only the DC source in effect, the capacitor appears to be an open circuit, and thus neither
it nor the shorted AC signal source will have any effect on the operation of the R 2 R3 voltage
divider:

4.8.

BIASING TECHNIQUES

225

2.3 V
R2

8.466
k
V1

R3

15 V

1.533
k

The capacitor appears to be an open circuit


as far as DC analysis is concerned
Combining these two separate analyses, we get a superposition of (almost) 1.5 volts AC and 2.3
volts DC, ready to be connected to the base of the transistor:

C
Vinput
1.5 V
2 kHz

R2

8.466
k
V1

R3

15 V

1.533
k

Enough talk its about time for a SPICE simulation of the whole amplifier circuit. Ill use a
capacitor value of 100 F to obtain an arbitrarily low (0.796 ) impedance at 2000 Hz:

voltage divider biasing


vinput 1 0 sin (0 1.5 2000 0 0)
c1 1 5 100u
r1 5 2 1k
r2 4 5 8466
r3 5 0 1533
q1 3 2 0 mod1
rspkr 3 4 8
v1 4 0 dc 15
.model mod1 npn
.tran 0.02m 0.78m
.plot tran v(1,0) i(v1)
.end

226

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

Notice that there is substantial distortion in the output waveform here: the sine wave is being
clipped during most of the input signals negative half-cycle. This tells us the transistor is entering
into cutoff mode when it shouldnt (Im assuming a goal of class A operation as before). Why is
this? This new biasing technique should give us exactly the same amount of DC bias voltage as
before, right?
With the capacitor and R2 R3 resistor network unloaded, it will provide exactly 2.3 volts
worth of DC bias. However, once we connect this network to the transistor, it is no longer unloaded.
Current drawn through the base of the transistor will load the voltage divider, thus reducing the
DC bias voltage available for the transistor. Using the diode-regulating diode transistor model to
illustrate, the bias problem becomes evident:
speaker
R2
C
Vinput

IR3 + Ibias
R1

R3

Ibias
IR3

Q1

V1

IR3

A voltage dividers output depends not only on the size of its constituent resistors, but also on
how much current is being divided away from it through a load. In this case, the base-emitter PN
junction of the transistor is a load that decreases the DC voltage dropped across R 3 , due to the fact
that the bias current joins with R3 s current to go through R2 , upsetting the divider ratio formerly
set by the resistance values of R2 and R3 . In order to obtain a DC bias voltage of 2.3 volts, the
values of R2 and/or R3 must be adjusted to compensate for the effect of base current loading. In

4.8.

BIASING TECHNIQUES

227

this case, we want to increase the DC voltage dropped across R3 , so we can lower the value of R2 ,
raise the value of R3 , or both.
voltage divider biasing
vinput 1 0 sin (0 1.5 2000 0 0)
c1 1 5 100u
r1 5 2 1k
r2 4 5 6k
<--- R2 decreased to 6 k ohms
r3 5 0 4k
<--- R3 increased to 4 k ohms
q1 3 2 0 mod1
rspkr 3 4 8
v1 4 0 dc 15
.model mod1 npn
.tran 0.02m 0.78m
.plot tran v(1,0) i(v1)
.end

As you can see, the new resistor values of 6 k and 4 k (R2 and R3 , respectively) results in
class A waveform reproduction, just the way we wanted.
REVIEW:
Class A operation is where an amplifier is biased so as to be in the active mode throughout
the entire waveform cycle, thus faithfully reproducing the whole waveform.
Class B operation is where an amplifier is biased so that only half of the input waveform gets
reproduced: either the positive half or the negative half. The transistor spends half its time
in the active mode and half its time cutoff. Complementary pairs of transistors running in

228

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS


class B operation are often used to deliver high power amplification in audio signal systems,
each transistor of the pair handling a separate half of the waveform cycle. Class B operation
delivers better power efficiency than a class A amplifier of similar output power.
Class AB operation is where an amplifier is biased at a point somewhere between class A and
class B.
Class C operation is where an amplifiers bias forces it to amplify only a small portion of the
waveform. A majority of the transistors time is spent in cutoff mode. In order for there to
be a complete waveform at the output, a resonant tank circuit is often used as a flywheel
to maintain oscillations for a few cycles after each kick from the amplifier. Because the
transistor is not conducting most of the time, power efficiencies are very high for a class C
amplifier.
Class D operation requires an advanced circuit design, and functions on the principle of representing instantaneous input signal amplitude by the duty cycle of a high-frequency squarewave.
The output transistor(s) never operate in active mode, only cutoff and saturation. Thus, there
is very little heat energy dissipated and energy efficiency is high.
DC bias voltage on the input signal, necessary for certain classes of operation (especially class
A and class C), may be obtained through the use of a voltage divider and coupling capacitor
rather than a battery connected in series with the AC signal source.

4.9

Input and output coupling

To overcome the challenge of creating necessary DC bias voltage for an amplifiers input signal
without resorting to the insertion of a battery in series with the AC signal source, we used a voltage
divider connected across the DC power source. To make this work in conjunction with an AC input
signal, we coupled the signal source to the divider through a capacitor, which acted as a high-pass
filter. With that filtering in place, the low impedance of the AC signal source couldnt short out
the DC voltage dropped across the bottom resistor of the voltage divider. A simple solution, but
not without any disadvantages.
Most obvious is the fact that using a high-pass filter capacitor to couple the signal source to
the amplifier means that the amplifier can only amplify AC signals. A steady, DC voltage applied
to the input would be blocked by the coupling capacitor just as much as the voltage divider bias
voltage is blocked from the input source. Furthermore, since capacitive reactance is frequencydependent, lower-frequency AC signals will not be amplified as much as higher-frequency signals.
Non-sinusoidal signals will tend to be distorted, as the capacitor responds differently to each of the
signals constituent harmonics. An extreme example of this would be a low-frequency square-wave
signal:

4.9.

INPUT AND OUTPUT COUPLING

229

R2

V1
Vinput

R3

Incidentally, this same problem occurs when oscilloscope inputs are set to the AC coupling
mode. In this mode, a coupling capacitor is inserted in series with the measured voltage signal to
eliminate any vertical offset of the displayed waveform due to DC voltage combined with the signal.
This works fine when the AC component of the measured signal is of a fairly high frequency, and the
capacitor offers little impedance to the signal. However, if the signal is of a low frequency, and/or
contains considerable levels of harmonics over a wide frequency range, the oscilloscopes display of
the waveform will not be accurate.

FUNCTION GENERATOR

40.00 Hz
coarse

fine

10 100 1k 10k 100k 1M

DC

output

OSCILLOSCOPE
vertical
Y
V/div

DC GND AC

trigger

timebase
X
s/div

DC GND AC

With DC coupling, the oscilloscope properly indicates the shape


of the square wave coming from the signal generator.

230

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS


FUNCTION GENERATOR

40.00 Hz
coarse

10 100 1k 10k 100k 1M

fine

DC

output

OSCILLOSCOPE
vertical
Y
DC GND AC

V/div
trigger

timebase
X
s/div

DC GND AC

With AC coupling, the high-pass filtering of the coupling capacitor


distorts the square waves shape so that what is seen is not
an accurate representation of the real voltage signal.

In applications where the limitations of capacitive coupling would be intolerable, another solution
may be used: direct coupling. Direct coupling avoids the use of capacitors or any other frequencydependent coupling component in favor of resistors. A direct-coupled amplifier circuit might look
something like this:

speaker
4
1
Vinput

Rinput

R2

3
2

Q1

4
V1

R3
0

With no capacitor to filter the input signal, this form of coupling exhibits no frequency dependence. DC and AC signals alike will be amplified by the transistor with the same gain (the transistor
itself may tend to amplify some frequencies better than others, but that is another subject entirely!).
If direct coupling works for DC as well as for AC signals, then why use capacitive coupling for
any application? One reason might be to avoid any unwanted DC bias voltage naturally present
in the signal to be amplified. Some AC signals may be superimposed on an uncontrolled DC
voltage right from the source, and an uncontrolled DC voltage would make reliable transistor biasing

4.9.

INPUT AND OUTPUT COUPLING

231

impossible. The high-pass filtering offered by a coupling capacitor would work well here to avoid
biasing problems.

Another reason to use capacitive coupling rather than direct is its relative lack of signal attenuation. Direct coupling through a resistor has the disadvantage of diminishing, or attenuating, the
input signal so that only a fraction of it reaches the base of the transistor. In many applications,
some attenuation is necessary anyway to prevent normal signal levels from overdriving the transistor into cutoff and saturation, so any attenuation inherent to the coupling network is useful anyway.
However, some applications require that there be no signal loss from the input connection to the
transistors base for maximum voltage gain, and a direct coupling scheme with a voltage divider for
bias simply wont suffice.

So far, weve discussed a couple of methods for coupling an input signal to an amplifier, but
havent addressed the issue of coupling an amplifiers output to a load. The example circuit used to
illustrate input coupling will serve well to illustrate the issues involved with output coupling.

In our example circuit, the load is a speaker. Most speakers are electromagnetic in design: that
is, they use the force generated by an lightweight electromagnet coil suspended within a strong
permanent-magnet field to move a thin paper or plastic cone, producing vibrations in the air which
our ears interpret as sound. An applied voltage of one polarity moves the cone outward, while a
voltage of the opposite polarity will move the cone inward. To exploit cones full freedom of motion,
the speaker must receive true (unbiased) AC voltage. DC bias applied to the speaker coil tends to
offset the cone from its natural center position, and this tends to limit the amount of back-and-forth
motion it can sustain from the applied AC voltage without overtraveling. However, our example
circuit applies a varying voltage of only one polarity across the speaker, because the speaker is
connected in series with the transistor which can only conduct current one way. This situation
would be unacceptable in the case of any high-power audio amplifier.

Somehow we need to isolate the speaker from the DC bias of the collector current so that it only
receives AC voltage. One way to achieve this goal is to couple the transistor collector circuit to the
speaker through a transformer:

232

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

speaker

R
Vinput

R
Q1

V1

Voltage induced in the secondary (speaker-side) of the transformer will be strictly due to variations in collector current, because the mutual inductance of a transformer only works on changes in
winding current. In other words, only the AC portion of the collector current signal will be coupled
to the secondary side for powering the speaker. The speaker will see true alternating current at
its terminals, without any DC bias.

Transformer output coupling works, and has the added benefit of being able to provide impedance
matching between the transistor circuit and the speaker coil with custom winding ratios. However,
transformers tend to be large and heavy, especially for high-power applications. Also, it is difficult
to engineer a transformer to handle signals over a wide range of frequencies, which is almost always
required for audio applications. To make matters worse, DC current through the primary winding
adds to the magnetization of the core in one polarity only, which tends to make the transformer
core saturate more easily in one AC polarity cycle than the other. This problem is reminiscent of
having the speaker directly connected in series with the transistor: a DC bias current tends to limit
how much output signal amplitude the system can handle without distortion. Generally, though, a
transformer can be designed to handle a lot more DC bias current than a speaker without running
into trouble, so transformer coupling is still a viable solution in most cases.

Another method to isolate the speaker from DC bias in the output signal is to alter the circuit
a bit and use a coupling capacitor in a manner similar to coupling the input signal to the amplifier:

4.9.

INPUT AND OUTPUT COUPLING

R
R
Vinput

233

C
speaker

V1

Q1
R

This circuit resembles the more conventional form of common-emitter amplifier, with the transistor collector connected to the battery through a resistor. The capacitor acts as a high-pass filter,
passing most of the AC voltage to the speaker while blocking all DC voltage. Again, the value
of this coupling capacitor is chosen so that its impedance at the expected signal frequency will be
arbitrarily low.
The blocking of DC voltage from an amplifiers output, be it via a transformer or a capacitor,
is useful not only in coupling an amplifier to a load, but also in coupling one amplifier to another
amplifier. Staged amplifiers are often used to achieve higher power gains than what would be
possible using a single transistor:
Three-stage common-emitter amplifier

Voutput

Vinput

First stage

Second stage

Third stage

While it is possible to directly couple each stage to the next (via a resistor rather than a capacitor), this makes the whole amplifier very sensitive to variations in the DC bias voltage of the first
stage, since that DC voltage will be amplified along with the AC signal until the last stage. In other

234

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

words, the biasing of the first stage will affect the biasing of the second stage, and so on. However,
if the stages are capacitively coupled as shown in the above illustration, the biasing of one stage
has no effect on the biasing of the next, because DC voltage is blocked from passing on to the next
stage.
Transformer coupling between amplifier stages is also a possibility, but less often seen due to
some of the problems inherent to transformers mentioned previously. One notable exception to this
rule is in the case of radio-frequency amplifiers where coupling transformers are typically small, have
air cores (making them immune to saturation effects), and can be made part of a resonant circuit
so as to block unwanted harmonic frequencies from passing on to subsequent stages. The use of
resonant circuits assumes that the signal frequency remains constant, of course, but this is typically
the case in radio circuitry. Also, the flywheel effect of LC tank circuits allows for class C operation
for high efficiency:
Three-stage tuned (RF) amplifier

Voutput

Vinput

First stage

Second stage

Third stage

Having said all this, it must be mentioned that it is possible to use direct coupling within a multistage transistor amplifier circuit. In cases where the amplifier is expected to handle DC signals, this
is the only alternative.
REVIEW:
Capacitive coupling acts like a high-pass filter on the input of an amplifier. This tends to make
the amplifiers voltage gain decrease at lower signal frequencies. Capacitive-coupled amplifiers
are all but unresponsive to DC input signals.
Direct coupling with a series resistor instead of a series capacitor avoids the problem of
frequency-dependent gain, but has the disadvantage of reducing amplifier gain for all signal
frequencies by attenuating the input signal.
Transformers and capacitors may be used to couple the output of an amplifier to a load, to
eliminate DC voltage from getting to the load.
Multi-stage amplifiers often make use of capacitive coupling between stages to eliminate problems with the bias from one stage affecting the bias of another.

4.10.

4.10

FEEDBACK

235

Feedback

If some percentage of an amplifiers output signal is connected to the input, so that the amplifier
amplifies part of its own output signal, we have what is known as feedback. Feedback comes in
two varieties: positive (also called regenerative), and negative (also called degenerative). Positive
feedback reinforces the direction of an amplifiers output voltage change, while negative feedback
does just the opposite.
A familiar example of feedback happens in public-address (PA) systems where someone holds
the microphone too close to a speaker: a high-pitched whine or howl ensues, because the audio
amplifier system is detecting and amplifying its own noise. Specifically, this is an example of positive
or regenerative feedback, as any sound detected by the microphone is amplified and turned into a
louder sound by the speaker, which is then detected by the microphone again, and so on . . . the
result being a noise of steadily increasing volume until the system becomes saturated and cannot
produce any more volume.
One might wonder what possible benefit feedback is to an amplifier circuit, given such an annoying
example as PA system howl. If we introduce positive, or regenerative, feedback into an amplifier
circuit, it has the tendency of creating and sustaining oscillations, the frequency of which determined
by the values of components handling the feedback signal from output to input. This is one way
to make an oscillator circuit to produce AC from a DC power supply. Oscillators are very useful
circuits, and so feedback has a definite, practical application for us.
Negative feedback, on the other hand, has a dampening effect on an amplifier: if the output
signal happens to increase in magnitude, the feedback signal introduces a decreasing influence into
the input of the amplifier, thus opposing the change in output signal. While positive feedback
drives an amplifier circuit toward a point of instability (oscillations), negative feedback drives it the
opposite direction: toward a point of stability.
An amplifier circuit equipped with some amount of negative feedback is not only more stable,
but it tends to distort the input waveform to a lesser degree and is generally capable of amplifying a
wider range of frequencies. The tradeoff for these advantages (there just has to be a disadvantage to
negative feedback, right?) is decreased gain. If a portion of an amplifiers output signal is fed back
to the input in such a way as to oppose any changes in the output, it will require a greater input
signal amplitude to drive the amplifiers output to the same amplitude as before. This constitutes
a decreased gain. However, the advantages of stability, lower distortion, and greater bandwidth are
worth the tradeoff in reduced gain for many applications.
Lets examine a simple amplifier circuit and see how we might introduce negative feedback into
it:

236

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

Rload

R1
R3

Voutput
Vinput

R2

The amplifier configuration shown here is a common-emitter, with a resistor bias network formed
by R1 and R2 . The capacitor couples Vinput to the amplifier so that the signal source doesnt have
a DC voltage imposed on it by the R1 /R2 divider network. Resistor R3 serves the purpose of
controlling voltage gain. We could omit if for maximum voltage gain, but since base resistors like
this are common in common-emitter amplifier circuits, well keep it in this schematic.
Like all common-emitter amplifiers, this one inverts the input signal as it is amplified. In other
words, a positive-going input voltage causes the output voltage to decrease, or go in the direction
of negative, and vice versa. If we were to examine the waveforms with oscilloscopes, it would look
something like this:

Rload

R1
R3

Vinput

R2

Because the output is an inverted, or mirror-image, reproduction of the input signal, any connection between the output (collector) wire and the input (base) wire of the transistor will result in
negative feedback:

4.10.

FEEDBACK

237

R1

Rfeedback

Rload

R3

Vinput

R2

The resistances of R1 , R2 , R3 , and Rf eedback function together as a signal-mixing network so


that the voltage seen at the base of the transistor (in reference to ground) is a weighted average of
the input voltage and the feedback voltage, resulting in signal of reduced amplitude going into the
transistor. As a result, the amplifier circuit will have reduced voltage gain, but improved linearity
(reduced distortion) and increased bandwidth.
A resistor connecting collector to base is not the only way to introduce negative feedback into this
amplifier circuit, though. Another method, although more difficult to understand at first, involves
the placement of a resistor between the transistors emitter terminal and circuit ground, like this:

A different method of introducing


negative feedback into the circuit

Rload

R1
R3

Vinput

R2

Rfeedback

This new feedback resistor drops voltage proportional to the emitter current through the transistor, and it does so in such a way as to oppose the input signals influence on the base-emitter junction
of the transistor. Lets take a closer look at the emitter-base junction and see what difference this
new resistor makes:

238

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

Icollector

Ibase
+

+
Vinput

VB-E

Iemitter

With no feedback resistor connecting the emitter to ground, whatever level of input signal
(Vinput ) makes it through the coupling capacitor and R1 /R2 /R3 resistor network will be impressed
directly across the base-emitter junction as the transistors input voltage (V BE ). In other words,
with no feedback resistor, VBE equals Vinput . Therefore, if Vinput increases by 100 mV, then
VBE likewise increases by 100 mV: a change in one is the same as a change in the other, since the
two voltages are equal to each other.
Now lets consider the effects of inserting a resistor (Rf eedback ) between the transistors emitter
lead and ground:

Icollector

Ibase
+

+
Vinput

VB-E
-

Iemitter

+
-

Rfeedback Vfeedback

Note how the voltage dropped across Rf eedback adds with VBE to equal Vinput . With Rf eedback
in the Vinput VBE loop, VBE will no longer be equal to Vinput . We know that Rf eedback
will drop a voltage proportional to emitter current, which is in turn controlled by the base current,
which is in turn controlled by the voltage dropped across the base-emitter junction of the transistor
(VBE ). Thus, if Vinput were to increase in a positive direction, it would increase VBE , causing
more base current, causing more collector (load) current, causing more emitter current, and causing
more feedback voltage to be dropped across Rf eedback . This increase of voltage drop across the
feedback resistor, though, subtracts from Vinput to reduce the VBE , so that the actual voltage
increase for VBE will be less than the voltage increase of Vinput . No longer will a 100 mV increase
in Vinput result in a full 100 mV increase for VBE , because the two voltages are not equal to each
other.
Consequently, the input voltage has less control over the transistor than before, and the voltage
gain for the amplifier is reduced: just what we expected from negative feedback.
In practical common-emitter circuits, negative feedback isnt just a luxury; its a necessity for
stable operation. In a perfect world, we could build and operate a common-emitter transistor
amplifier with no negative feedback, and have the full amplitude of Vinput impressed across the

4.10.

FEEDBACK

239

transistors base-emitter junction. This would give us a large voltage gain. Unfortunately, though,
the relationship between base-emitter voltage and base-emitter current changes with temperature,
as predicted by the diode equation. As the transistor heats up, there will be less of a forward
voltage drop across the base-emitter junction for any given current. This causes a problem for us, as
the R1 /R2 voltage divider network is designed to provide the correct quiescent current through the
base of the transistor so that it will operate in whatever class of operation we desire (in this example,
Ive shown the amplifier working in class-A mode). If the transistors voltage/current relationship
changes with temperature, the amount of DC bias voltage necessary for the desired class of operation
will change. In this case, a hot transistor will draw more bias current for the same amount of bias
voltage, making it heat up even more, drawing even more bias current. The result, if unchecked, is
called thermal runaway.
Common-collector amplifiers, however, do not suffer from thermal runaway. Why is this? The
answer has everything to do with negative feedback:

A common-collector amplifier

R1
R3

Vinput

R2

Rload

Note that the common-collector amplifier has its load resistor placed in exactly the same spot
as we had the Rf eedback resistor in the last circuit: between emitter and ground. This means that
the only voltage impressed across the transistors base-emitter junction is the difference between
Vinput and Voutput , resulting in a very low voltage gain (usually close to 1 for a common-collector
amplifier). Thermal runaway is impossible for this amplifier: if base current happens to increase
due to transistor heating, emitter current will likewise increase, dropping more voltage across the
load, which in turn subtracts from Vinput to reduce the amount of voltage dropped between base
and emitter. In other words, the negative feedback afforded by placement of the load resistor makes
the problem of thermal runaway self-correcting. In exchange for a greatly reduced voltage gain, we
get superb stability and immunity from thermal runaway.
By adding a feedback resistor between emitter and ground in a common-emitter amplifier,
we make the amplifier behave a little less like an ideal common-emitter and a little more like a
common-collector. The feedback resistor value is typically quite a bit less than the load, minimizing
the amount of negative feedback and keeping the voltage gain fairly high.
Another benefit of negative feedback, seen clearly in the common-collector circuit, is that it tends
to make the voltage gain of the amplifier less dependent on the characteristics of the transistor.

240

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

Note that in a common-collector amplifier, voltage gain is nearly equal to unity (1), regardless of the
transistors . This means, among other things, that we could replace the transistor in a commoncollector amplifier with one having a different and not see any significant changes in voltage gain.
In a common-emitter circuit, the voltage gain is highly dependent on . If we were to replace the
transistor in a common-emitter circuit with another of differing , the voltage gain for the amplifier
would change significantly. In a common-emitter amplifier equipped with negative feedback, the
voltage gain will still be dependent upon transistor to some degree, but not as much as before,
making the circuit more predictable despite variations in transistor .
The fact that we have to introduce negative feedback into a common-emitter amplifier to avoid
thermal runaway is an unsatisfying solution. It would be nice, after all, to avoid thermal runaway
without having to suppress the amplifiers inherently high voltage gain. A best-of-both-worlds
solution to this dilemma is available to us if we closely examine the nature of the problem: the
voltage gain that we have to minimize in order to avoid thermal runaway is the DC voltage gain,
not the AC voltage gain. After all, it isnt the AC input signal that fuels thermal runaway: its the
DC bias voltage required for a certain class of operation: that quiescent DC signal that we use to
trick the transistor (fundamentally a DC device) into amplifying an AC signal. We can suppress
DC voltage gain in a common-emitter amplifier circuit without suppressing AC voltage gain if we
figure out a way to make the negative feedback function with DC only. That is, if we only feed back
an inverted DC signal from output to input, but not an inverted AC signal.
The Rf eedback emitter resistor provides negative feedback by dropping a voltage proportional to
load current. In other words, negative feedback is accomplished by inserting an impedance into the
emitter current path. If we want to feed back DC but not AC, we need an impedance that is high
for DC but low for AC. What kind of circuit presents a high impedance to DC but a low impedance
to AC? A high-pass filter, of course!
By connecting a capacitor in parallel with the feedback resistor, we create the very situation we
need: a path from emitter to ground that is easier for AC than it is for DC:

High AC voltage gain re-established


by adding Cbypass in parallel with Rfeedback

Rload

R1
R3

Vinput

R2

Rfeedback

Cbypass

The new capacitor bypasses AC from the transistors emitter to ground, so that no appreciable
AC voltage will be dropped from emitter to ground to feed back to the input and suppress voltage
gain. Direct current, on the other hand, cannot go through the bypass capacitor, and so must travel

4.10.

FEEDBACK

241

through the feedback resistor, dropping a DC voltage between emitter and ground which lowers the
DC voltage gain and stabilizes the amplifiers DC response, preventing thermal runaway. Because we
want the reactance of this capacitor (XC ) to be as low as possible, Cbypass should be sized relatively
large. Because the polarity across this capacitor will never change, it is safe to use a polarized
(electrolytic) capacitor for the task.
Another approach to the problem of negative feedback reducing voltage gain is to use multistage amplifiers rather than single-transistor amplifiers. If the attenuated gain of a single transistor
is insufficient for the task at hand, we can use more than one transistor to make up for the reduction
caused by feedback. Here is an example circuit showing negative feedback in a three-stage commonemitter amplifier:

Rfeedback

Rin
Vinput

Voutput

Note how there is but one path for feedback, from the final output to the input through a single
resistor, Rf eedback . Since each stage is a common-emitter amplifier and thus inverting in nature
and there are an odd number of stages from input to output, the output signal will be inverted
with respect to the input signal, and the feedback will be negative (degenerative). Relatively large
amounts of feedback may be used without sacrificing voltage gain, because the three amplifier stages
provide so much gain to begin with.
At first, this design philosophy may seem inelegant and perhaps even counter-productive. Isnt
this a rather crude way to overcome the loss in gain incurred through the use of negative feedback,
to simply recover gain by adding stage after stage? What is the point of creating a huge voltage
gain using three transistor stages if were just going to attenuate all that gain anyway with negative
feedback? The point, though perhaps not apparent at first, is increased predictability and stability
from the circuit as a whole. If the three transistor stages are designed to provide an arbitrarily high
voltage gain (in the tens of thousands, or greater) with no feedback, it will be found that the addition
of negative feedback causes the overall voltage gain to become less dependent of the individual stage
gains, and approximately equal to the simple ratio Rf eedback /Rin . The more voltage gain the circuit
has (without feedback), the more closely the voltage gain will approximate R f eedback /Rin once
feedback is established. In other words, voltage gain in this circuit is fixed by the values of two
resistors, and nothing more.
This advantage has profound impact on mass-production of electronic circuitry: if amplifiers
of predictable gain may be constructed using transistors of widely varied values, it makes the
selection and replacement of components very easy and inexpensive. It also means the amplifiers
gain varies little with changes in temperature. This principle of stable gain control through a highgain amplifier tamed by negative feedback is elevated almost to an art form in electronic circuits

242

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

called operational amplifiers, or op-amps. You may read much more about these circuits in a later
chapter of this book!
REVIEW:
Feedback is the coupling of an amplifiers output to its input.
Positive, or regenerative feedback has the tendency of making an amplifier circuit unstable, so
that it produces oscillations (AC). The frequency of these oscillations is largely determined by
the components in the feedback network.
Negative, or degenerative feedback has the tendency of making an amplifier circuit more stable,
so that its output changes less for a given input signal than without feedback. This reduces the
gain of the amplifier, but has the advantage of decreasing distortion and increasing bandwidth
(the range of frequencies the amplifier can handle).
Negative feedback may be introduced into a common-emitter circuit by coupling collector to
base, or by inserting a resistor between emitter and ground.
An emitter-to-ground feedback resistor is usually found in common-emitter circuits as a
preventative measure against thermal runaway.
Negative feedback also has the advantage of making amplifier voltage gain more dependent on
resistor values and less dependent on the transistors characteristics.
Common-collector amplifiers have a lot of negative feedback, due to the placement of the load
resistor between emitter and ground. This feedback accounts for the extremely stable voltage
gain of the amplifier, as well as its immunity against thermal runaway.
Voltage gain for a common-emitter circuit may be re-established without sacrificing immunity
to thermal runaway, by connecting a bypass capacitor in parallel with the emitter feedback
resistor.
If the voltage gain of an amplifier is arbitrarily high (tens of thousands, or greater), and
negative feedback is used to reduce the gain to reasonable levels, it will be found that the gain
will approximately equal Rf eedback /Rin . Changes in transistor or other internal component
values will have comparatively little effect on voltage gain with feedback in operation, resulting
in an amplifier that is stable and easy to design.

4.11

Amplifier impedances

*** PENDING ***


REVIEW:

4.12.

CURRENT MIRRORS

4.12

243

Current mirrors

An interesting and often-used circuit applying the bipolar junction transistor is the so-called current
mirror, which serves as a simple current regulator, supplying nearly constant current to a load over
a wide range of load resistances.
We know that in a transistor operating in its active mode, collector current is equal to base
current multiplied by the ratio . We also know that the ratio between collector current and emitter
current is called . Because collector current is equal to base current multiplied by , and emitter
current is the sum of the base and collector currents, should be mathematically derivable from .
If you do the algebra, youll find that = /(+1) for any transistor.
Weve seen already how maintaining a constant base current through an active transistor results
in the regulation of collector current, according to the ratio. Well, the ratio works similarly: if
emitter current is held constant, collector current will remain at a stable, regulated value so long
as the transistor has enough collector-to-emitter voltage drop to maintain it in its active mode.
Therefore, if we have a way of holding emitter current constant through a transistor, the transistor
will work to regulate collector current at a constant value.
Remember that the base-emitter junction of a BJT is nothing more than a PN junction, just like
a diode, and that the diode equation specifies how much current will go through a PN junction
given forward voltage drop and junction temperature:

ID = IS (eqVD/NkT - 1)
Where,
ID = Diode current in amps
IS = Saturation current in amps
(typically 1 x 10-12 amps)
e = Eulers constant (~ 2.718281828)
q = charge of electron (1.6 x 10-19 coulombs)
VD = Voltage applied across diode in volts
N = "Nonideality" or "emission" coefficient
(typically between 1 and 2)
k = Boltzmanns constant (1.38 x 10-23)
T = Junction temperature in Kelvins
If both junction voltage and temperature are held constant, then the PN junction current will
likewise be constant. Following this rationale, if we were to hold the base-emitter voltage of a transistor constant, then its emitter current should likewise be constant, given a constant temperature:

244

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

Icollector

Rload

(constant)

Ibase
Vbase

(constant)

(constant)
(constant)
Iemitter

(constant)

(constant)

This constant emitter current, multiplied by a constant ratio, gives a constant collector current
through Rload , provided that there is enough battery voltage to keep the transistor in its active mode
for any change in Rload s resistance.
Maintaining a constant voltage across the transistors base-emitter junction is easy: use a
forward-biased diode to establish a constant voltage of approximately 0.7 volts, and connect it
in parallel with the base-emitter junction:

Rbias

Icollector
(constant)

Ibase (constant)
0.7 V
(constant)

Idiode
(constant)

Rload
(constant)
(constant)
Iemitter
(constant)

Now, heres where it gets interesting. The voltage dropped across the diode probably wont be
0.7 volts exactly. The exact amount of forward voltage dropped across it depends on the current
through the diode, and the diodes temperature, all in accordance with the diode equation. If
diode current is increased (say, by reducing the resistance of Rbias ), its voltage drop will increase
slightly, increasing the voltage drop across the transistors base-emitter junction, which will increase
the emitter current by the same proportion, assuming the diodes PN junction and the transistors
base-emitter junction are well-matched to each other. In other words, transistor emitter current
will closely equal diode current at any given time. If you change the diode current by changing the
resistance value of Rbias , then the transistors emitter current will follow suit, because the emitter
current is described by the same equation as the diodes, and both PN junctions experience the same
voltage drop.
Remember, the transistors collector current is almost equal to its emitter current, as the ratio
of a typical transistor is almost unity (1). If we have control over the transistors emitter current
by setting diode current with a simple resistor adjustment, then we likewise have control over the
transistors collector current. In other words, collector current mimics, or mirrors, diode current.
Current through resistor Rload is therefore a function of current set by the bias resistor, the two

4.12.

CURRENT MIRRORS

245

being nearly equal. This is the function of the current mirror circuit: to regulate current through the
load resistor by conveniently adjusting the value of Rbias . It is very easy to create a set amount of
diode current, as current through the diode is described by a simple equation: power supply voltage
minus diode voltage (almost a constant value), divided by the resistance of R bias .
To better match the characteristics of the two PN junctions (the diode junction and the transistor
base-emitter junction), a transistor may be used in place of a regular diode, like this:

A current mirror circuit


using two transistors

Rbias

Rload

Because temperature is a factor in the diode equation, and we want the two PN junctions to
behave identically under all operating conditions, we should maintain the two transistors at exactly
the same temperature. This is easily done using discrete components by gluing the two transistor
cases back-to-back. If the transistors are manufactured together on a single chip of silicon (as a
so-called integrated circuit, or IC ), the designers should locate the two transistors very close to one
another to facilitate heat transfer between them.
The current mirror circuit shown with two NPN transistors is sometimes called a current-sinking
type, because the regulating transistor conducts current to the load from ground (sinking current),
rather than from the positive side of the battery (sourcing current). If we wish to have a grounded
load, and a current sourcing mirror circuit, we could use PNP transistors like this:

A current-sourcing mirror circuit

Rbias

REVIEW:

Rload

246

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS


A current mirror is a transistor circuit that regulates current through a load resistance, the
regulation point being set by a simple resistor adjustment.
Transistors in a current mirror circuit must be maintained at the same temperature for precise operation. When using discrete transistors, you may glue their cases together to help
accomplish this.
Current mirror circuits may be found in two basic varieties: the current sinking configuration,
where the regulating transistor connects the load to ground; and the current sourcing configuration, where the regulating transistor connects the load to the positive terminal of the DC
power supply.

4.13

Transistor ratings and packages

*** INCOMPLETE ***


Like all electrical and electronic components, transistors are limited in the amounts of voltage
and current they can handle without sustaining damage. Since transistors are a bit more complex
than some of the other components youre used to seeing at this point, they tend to have more kinds
of ratings. What follows is an itemized description of some typical transistor ratings.
Power dissipation: When a transistor conducts current between collector and emitter, it also
drops voltage between those two points. At any given time, the power dissipated by a transistor
is equal to the product (multiplication) of collector current and collector-emitter voltage. Just
like resistors, transistors are rated in terms of how many watts they can safely dissipate without
sustaining damage. High temperature is the mortal enemy of all semiconductor devices, and bipolar
transistors tend to be more susceptible to thermal damage than most. Power ratings are always
given in reference to the temperature of ambient (surrounding) air. When transistors are to be
used in hotter-than-normal environments, their power ratings must be derated to avoid a shortened
service life.
Reverse voltages: As with diodes, bipolar transistors are rated for maximum allowable reversebias voltage across their PN junctions. This includes voltage ratings for the base-emitter junction,
base-collector junction, and also from collector to emitter. The rating for maximum collector-emitter
voltage can be thought of in terms of the maximum voltage it can withstand while in full-cutoff mode
(no base current). This rating is of particular importance when using a bipolar transistor as a switch.
Collector current: A maximum value for collector current will be given by the manufacturer in
amps. Understand that this maximum figure assumes a saturated state (minimum collector-emitter
voltage drop). If the transistor is not saturated, and in fact is dropping substantial voltage between
collector and emitter, the maximum power dissipation rating will probably be exceeded before the
maximum collector current rating will. Just something to keep in mind when designing a transistor
circuit!
Saturation voltages: Ideally, a saturated transistor acts as a closed switch contact between collector and emitter, dropping zero voltage at full collector current. In reality this is never true.
Manufacturers will specify the maximum voltage drop of a transistor at saturation, both between
the collector and emitter, and also between base and emitter (forward voltage drop of that PN
junction). Collector-emitter voltage drop at saturation is generally expected to be 0.3 volts or less,
but this figure is of course dependent on the specific type of transistor. Base-emitter forward voltage
drop is very similar to that of an equivalent diode, which should come as no surprise.

4.14.

BJT QUIRKS

247

Beta: The ratio of collector current to base current, is the fundamental parameter characterizing the amplifying ability of a bipolar transistor. is usually assumed to be a constant figure
in circuit calculations, but unfortunately this is far from true in practice. As such, manufacturers
provide a set of (or hf e ) figures for a given transistor over a wide range of operating conditions,
usually in the form of maximum/minimum/typical ratings. It may surprise you to see just how
widely can be expected to vary within normal operating limits. One popular small-signal transistor, the 2N3903, is advertised as having a ranging from 15 to 150 depending on the amount of
collector current. Generally, is highest for medium collector currents, decreasing for very low and
very high collector currents.
Alpha: the ratio of collector current to emitter current, may be derived from , being equal to
/(+1).
Bipolar transistors come in a wide variety of physical packages. Package type is primarily dependent upon the power dissipation of the transistor, much like resistors: the greater the maximum
power dissipation, the larger the device has to be to stay cool. There are several standardized package types for three-terminal semiconductor devices, any of which may be used to house a bipolar
transistor. This is an important fact to consider: there are many other semiconductor devices other
than bipolar transistors which have three connection points. It is impossible to positively identify
a three-terminal semiconductor device without referencing the part number printed on it, and/or
subjecting it to a set of electrical tests.
REVIEW:

4.14

BJT quirks

*** PENDING ***


Nonlinearity Temperature drift Thermal runaway Junction capacitance Noise Mismatch (problem
with paralleling transistors) cutoff frequency Alpha cutoff frequency
REVIEW:

248

CHAPTER 4. BIPOLAR JUNCTION TRANSISTORS

Chapter 5

JUNCTION FIELD-EFFECT
TRANSISTORS
Contents
5.1
5.2
5.3
5.4
5.5
5.6
5.7
5.8
5.9
5.10

Introduction . . . . . . . . . . . . . . . . . . . .
The transistor as a switch . . . . . . . . . . . .
Meter check of a transistor . . . . . . . . . . .
Active-mode operation . . . . . . . . . . . . . .
The common-source amplifier PENDING .
The common-drain amplifier PENDING . .
The common-gate amplifier PENDING . .
Biasing techniques PENDING . . . . . . . .
Transistor ratings and packages PENDING
JFET quirks PENDING . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

249
251
254
256
265
266
266
266
266
267

*** INCOMPLETE ***

5.1

Introduction

A transistor is a linear semiconductor device that controls current with the application of a lowerpower electrical signal. Transistors may be roughly grouped into two major divisions: bipolar and
field-effect. In the last chapter we studied bipolar transistors, which utilize a small current to control
a large current. In this chapter, well introduce the general concept of the field-effect transistor
a device utilizing a small voltage to control current and then focus on one particular type: the
junction field-effect transistor. In the next chapter well explore another type of field-effect transistor,
the insulated gate variety.
All field-effect transistors are unipolar rather than bipolar devices. That is, the main current
through them is comprised either of electrons through an N-type semiconductor or holes through a
P-type semiconductor. This becomes more evident when a physical diagram of the device is seen:
249

250

CHAPTER 5. JUNCTION FIELD-EFFECT TRANSISTORS

N-channel JFET
drain
drain
gate

gate

P N

source
source
schematic symbol

physical diagram

In a junction field-effect transistor, or JFET, the controlled current passes from source to drain,
or from drain to source as the case may be. The controlling voltage is applied between the gate
and source. Note how the current does not have to cross through a PN junction on its way between
source and drain: the path (called a channel ) is an uninterrupted block of semiconductor material.
In the image just shown, this channel is an N-type semiconductor. P-type channel JFETs are also
manufactured:

P-channel JFET
drain
drain
gate

gate

N P

source
source
schematic symbol

physical diagram

Generally, N-channel JFETs are more commonly used than P-channel. The reasons for this have
to do with obscure details of semiconductor theory, which Id rather not discuss in this chapter. As
with bipolar transistors, I believe the best way to introduce field-effect transistor usage is to avoid
theory whenever possible and concentrate instead on operational characteristics. The only practical
difference between N- and P-channel JFETs you need to concern yourself with now is biasing of the
PN junction formed between the gate material and the channel.
With no voltage applied between gate and source, the channel is a wide-open path for electrons to
flow. However, if a voltage is applied between gate and source of such polarity that it reverse-biases
the PN junction, the flow between source and drain connections becomes limited, or regulated, just
as it was for bipolar transistors with a set amount of base current. Maximum gate-source voltage

5.2.

THE TRANSISTOR AS A SWITCH

251

pinches off all current through source and drain, thus forcing the JFET into cutoff mode. This
behavior is due to the depletion region of the PN junction expanding under the influence of a reversebias voltage, eventually occupying the entire width of the channel if the voltage is great enough.
This action may be likened to reducing the flow of a liquid through a flexible hose by squeezing it:
with enough force, the hose will be constricted enough to completely block the flow.

water

hose

nozzle

water

Hose constricted by squeezing,


water flow reduced or stopped
Note how this operational behavior is exactly opposite of the bipolar junction transistor. Bipolar
transistors are normally-off devices: no current through the base, no current through the collector
or the emitter. JFETs, on the other hand, are normally-on devices: no voltage applied to the gate
allows maximum current through the source and drain. Also take note that the amount of current
allowed through a JFET is determined by a voltage signal rather than a current signal as with bipolar
transistors. In fact, with the gate-source PN junction reverse-biased, there should be nearly zero
current through the gate connection. For this reason, we classify the JFET as a voltage-controlled
device, and the bipolar transistor as a current-controlled device.
If the gate-source PN junction is forward-biased with a small voltage, the JFET channel will
open a little more to allow greater currents through. However, the PN junction of a JFET is not
built to handle any substantial current itself, and thus it is not recommended to forward-bias the
junction under any circumstances.
This is a very condensed overview of JFET operation. In the next section, well explore the use
of the JFET as a switching device.

5.2

The transistor as a switch

Like its bipolar cousin, the field-effect transistor may be used as an on/off switch controlling electrical
power to a load. Lets begin our investigation of the JFET as a switch with our familiar switch/lamp
circuit:

switch

252

CHAPTER 5. JUNCTION FIELD-EFFECT TRANSISTORS

Remembering that the controlled current in a JFET flows between source and drain, we substitute
the source and drain connections of a JFET for the two ends of the switch in the above circuit:

If you havent noticed by now, the source and drain connections on a JFET look identical on the
schematic symbol. Unlike the bipolar junction transistor where the emitter is clearly distinguished
from the collector by the arrowhead, a JFETs source and drain lines both run perpendicular into
the bar representing the semiconductor channel. This is no accident, as the source and drain lines
of a JFET are often interchangeable in practice! In other words, JFETs are usually able to handle
channel current in either direction, from source to drain or from drain to source.
Now all we need in the circuit is a way to control the JFETs conduction. With zero applied
voltage between gate and source, the JFETs channel will be open, allowing full current to the
lamp. In order to turn the lamp off, we will need to connect another source of DC voltage between
the gate and source connections of the JFET like this:

switch

Closing this switch will pinch off the JFETs channel, thus forcing it into cutoff and turning
the lamp off:

switch

Note that there is no current going through the gate. As a reverse-biased PN junction, it firmly
opposes the flow of any electrons through it. As a voltage-controlled device, the JFET requires
negligible input current. This is an advantageous trait of the JFET over the bipolar transistor:
there is virtually zero power required of the controlling signal.
Opening the control switch again should disconnect the reverse-biasing DC voltage from the gate,
thus allowing the transistor to turn back on. Ideally, anyway, this is how it works. In practice this
may not work at all:

5.2.

THE TRANSISTOR AS A SWITCH

253

switch

No lamp current after the switch opens!


Why is this? Why doesnt the JFETs channel open up again and allow lamp current through
like it did before with no voltage applied between gate and source? The answer lies in the operation
of the reverse-biased gate-source junction. The depletion region within that junction acts as an
insulating barrier separating gate from source. As such, it possesses a certain amount of capacitance
capable of storing an electric charge potential. After this junction has been forcibly reverse-biased
by the application of an external voltage, it will tend to hold that reverse-biasing voltage as a stored
charge even after the source of that voltage has been disconnected. What is needed to turn the
JFET on again is to bleed off that stored charge between the gate and source through a resistor:

switch

Resistor bleeds off stored charge in


PN junction to allow transistor to
turn on once again.
This resistors value is not very important. The capacitance of the JFETs gate-source junction
is very small, and so even a rather high-value bleed resistor creates a fast RC time constant, allowing
the transistor to resume conduction with little delay once the switch is opened.
Like the bipolar transistor, it matters little where or what the controlling voltage comes from.
We could use a solar cell, thermocouple, or any other sort of voltage-generating device to supply the
voltage controlling the JFETs conduction. All that is required of a voltage source for JFET switch
operation is sufficient voltage to achieve pinch-off of the JFET channel. This level is usually in the
realm of a few volts DC, and is termed the pinch-off or cutoff voltage. The exact pinch-off voltage
for any given JFET is a function of its unique design, and is not a universal figure like 0.7 volts is
for a silicon BJTs base-emitter junction voltage.
REVIEW:
Field-effect transistors control the current between source and drain connections by a voltage
applied between the gate and source. In a junction field-effect transistor (JFET), there is
a PN junction between the gate and source which is normally reverse-biased for control of
source-drain current.
JFETs are normally-on (normally-saturated) devices. The application of a reverse-biasing

254

CHAPTER 5. JUNCTION FIELD-EFFECT TRANSISTORS


voltage between gate and source causes the depletion region of that junction to expand, thereby
pinching off the channel between source and drain through which the controlled current
travels.

It may be necessary to attach a bleed-off resistor between gate and source to discharge the
stored charge built up across the junctions natural capacitance when the controlling voltage
is removed. Otherwise, a charge may remain to keep the JFET in cutoff mode even after the
voltage source has been disconnected.

5.3

Meter check of a transistor

Testing a JFET with a multimeter might seem to be a relatively easy task, seeing as how it has only
one PN junction to test: either measured between gate and source, or between gate and drain.

N-channel transistor

OFF

drain
+

COM

drain
gate

gate

P N

source
+
source
V

A
physical diagram

A
OFF

COM

Both meters show non-continuity


(high resistance) through gatechannel junction.

5.3.

METER CHECK OF A TRANSISTOR

255

N-channel transistor

OFF

drain
-

COM

drain
gate

gate

+
+

P N

source
source
V

physical diagram
OFF

COM

Both meters show continuity (low


resistance) through gate-channel
junction.

Testing continuity through the drain-source channel is another matter, though. Remember from
the last section how a stored charge across the capacitance of the gate-channel PN junction could
hold the JFET in a pinched-off state without any external voltage being applied across it? This
can occur even when youre holding the JFET in your hand to test it! Consequently, any meter
reading of continuity through that channel will be unpredictable, since you dont necessarily know
if a charge is being stored by the gate-channel junction. Of course, if you know beforehand which
terminals on the device are the gate, source, and drain, you may connect a jumper wire between
gate and source to eliminate any stored charge and then proceed to test source-drain continuity
with no problem. However, if you dont know which terminals are which, the unpredictability of the
source-drain connection may confuse your determination of terminal identity.
A good strategy to follow when testing a JFET is to insert the pins of the transistor into antistatic foam (the material used to ship and store static-sensitive electronic components) just prior to
testing. The conductivity of the foam will make a resistive connection between all terminals of the
transistor when it is inserted. This connection will ensure that all residual voltage built up across
the gate-channel PN junction will be neutralized, thus opening up the channel for an accurate
meter test of source-to-drain continuity.
Since the JFET channel is a single, uninterrupted piece of semiconductor material, there is usually
no difference between the source and drain terminals. A resistance check from source to drain should
yield the same value as a check from drain to source. This resistance should be relatively low (a few
hundred ohms at most) when the gate-source PN junction voltage is zero. By applying a reversebias voltage between gate and source, pinch-off of the channel should be apparent by an increased
resistance reading on the meter.

256

CHAPTER 5. JUNCTION FIELD-EFFECT TRANSISTORS

5.4

Active-mode operation

JFETs, like bipolar transistors, are able to throttle current in a mode between cutoff and saturation
called the active mode. To better understand JFET operation, lets set up a SPICE simulation
similar to the one used to explore basic bipolar transistor function:

Vammeter
2

3
0V

V1

Q1

Vin

jfet simulation
vin 0 1 dc 1
j1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc
.model mod1 njf
.dc v1 0 2 0.05
.plot dc i(vammeter)
.end
Note that the transistor labeled Q1 in the schematic is represented in the SPICE netlist as
j1. Although all transistor types are commonly referred to as Q devices in circuit schematics
just as resistors are referred to by R designations, and capacitors by C SPICE needs to be
told what type of transistor this is by means of a different letter designation: q for bipolar junction
transistors, and j for junction field-effect transistors.

5.4.

ACTIVE-MODE OPERATION

257

Here, the controlling signal is a steady voltage of 1 volt, applied with negative towards the
JFET gate and positive toward the JFET source, to reverse-bias the PN junction. In the first BJT
simulation of chapter 4, a constant-current source of 20 A was used for the controlling signal, but
remember that a JFET is a voltage-controlled device, not a current-controlled device like the bipolar
junction transistor.
Like the BJT, the JFET tends to regulate the controlled current at a fixed level above a certain
power supply voltage, no matter how high that voltage may climb. Of course, this current regulation
has limits in real life no transistor can withstand infinite voltage from a power source and with
enough drain-to-source voltage the transistor will break down and drain current will surge. But
within normal operating limits the JFET keeps the drain current at a steady level independent of
power supply voltage. To verify this, well run another computer simulation, this time sweeping the
power supply voltage (V1 ) all the way to 50 volts:

jfet simulation
vin 0 1 dc 1
j1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc
.model mod1 njf
.dc v1 0 50 2
.plot dc i(vammeter)
.end

258

CHAPTER 5. JUNCTION FIELD-EFFECT TRANSISTORS

Sure enough, the drain current remains steady at a value of 100 A (1.000E-04 amps) no matter
how high the power supply voltage is adjusted.
Because the input voltage has control over the constriction of the JFETs channel, it makes sense
that changing this voltage should be the only action capable of altering the current regulation point
for the JFET, just like changing the base current on a BJT is the only action capable of altering
collector current regulation. Lets decrease the input voltage from 1 volt to 0.5 volts and see what
happens:

jfet simulation
vin 0 1 dc 0.5
j1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc
.model mod1 njf
.dc v1 0 50 2
.plot dc i(vammeter)
.end

5.4.

ACTIVE-MODE OPERATION

259

As expected, the drain current is greater now than it was in the previous simulation. With less
reverse-bias voltage impressed across the gate-source junction, the depletion region is not as wide as
it was before, thus opening the channel for charge carriers and increasing the drain current figure.
Please note, however, the actual value of this new current figure: 225 A (2.250E-04 amps). The
last simulation showed a drain current of 100 A, and that was with a gate-source voltage of 1 volt.
Now that weve reduced the controlling voltage by a factor of 2 (from 1 volt down to 0.5 volts), the
drain current increased, but not by the same 2:1 proportion! Lets reduce our gate-source voltage
once more by another factor of 2 (down to 0.25 volts) and see what happens:

jfet simulation
vin 0 1 dc 0.25
j1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc
.model mod1 njf
.dc v1 0 50 2
.plot dc i(vammeter)
.end

260

CHAPTER 5. JUNCTION FIELD-EFFECT TRANSISTORS

With the gate-source voltage set to 0.25 volts, one-half what it was before, the drain current
is 306.3 A. Although this is still an increase over the 225 A from the prior simulation, it isnt
proportional to the change of the controlling voltage.
To obtain a better understanding of what is going on here, we should run a different kind of
simulation: one that keeps the power supply voltage constant and instead varies the controlling
(voltage) signal. When this kind of simulation was run on a BJT, the result was a straight-line
graph, showing how the input current / output current relationship of a BJT is linear. Lets see
what kind of relationship a JFET exhibits:

jfet simulation
vin 0 1 dc
j1 2 1 0 mod1
vammeter 3 2 dc 0
v1 3 0 dc 25
.model mod1 njf
.dc vin 0 2 0.1
.plot dc i(vammeter)
.end

5.4.

ACTIVE-MODE OPERATION

261

This simulation directly reveals an important characteristic of the junction field-effect transistor:
the control effect of gate voltage over drain current is nonlinear. Notice how the drain current does
not decrease linearly as the gate-source voltage is increased. With the bipolar junction transistor,
collector current was directly proportional to base current: output signal proportionately followed
input signal. Not so with the JFET! The controlling signal (gate-source voltage) has less and less
effect over the drain current as it approaches cutoff. In this simulation, most of the controlling action
(75 percent of drain current decrease from 400 A to 100 A) takes place within the first volt of
gate-source voltage (from 0 to 1 volt), while the remaining 25 percent of drain current reduction
takes another whole volt worth of input signal. Cutoff occurs at 2 volts input.
Linearity is generally important for a transistor because it allows it to faithfully amplify a waveform without distorting it. If a transistor is nonlinear in its input/output amplification, the shape of
the input waveform will become corrupted in some way, leading to the production of harmonics in
the output signal. The only time linearity is not important in a transistor circuit is when its being
operated at the extreme limits of cutoff and saturation (off and on, respectively, like a switch).
A JFETs characteristic curves display the same current-regulating behavior as for a BJT, and
the nonlinearity between gate-to-source voltage and drain current is evident in the disproportionate
vertical spacings between the curves:

262

CHAPTER 5. JUNCTION FIELD-EFFECT TRANSISTORS

|VDS| = |VP| - |VGS|


Below pinch-off
Triode region

Above pinch-off
Saturation region

Idrain

Ohmic
regio
n

Vgate-to-source = 0 V

Vgate-to-source = 0.5 V

Vgate-to-source = 1 V
Vgate-to-source = 2 V = VP

(pinch-off)

Edrain-to-source
To better comprehend the current-regulating behavior of the JFET, it might be helpful to draw
a model made up of simpler, more common components, just as we did for the BJT:

D
G
S
N-channel JFET diode-regulating diode model
D

S
In the case of the JFET, it is the voltage across the reverse-biased gate-source diode which sets
the current regulation point for the pair of constant-current diodes. A pair of opposing constantcurrent diodes is included in the model to facilitate current in either direction between source and

5.4.

ACTIVE-MODE OPERATION

263

drain, a trait made possible by the unipolar nature of the channel. With no PN junctions for the
source-drain current to traverse, there is no polarity sensitivity in the controlled current. For this
reason, JFETs are often referred to as bilateral devices.
A contrast of the JFETs characteristic curves against the curves for a bipolar transistor reveals
a notable difference: the linear (straight) portion of each curves non-horizontal area is surprisingly
long compared to the respective portions of a BJTs characteristic curves:

Vgate-to-source = 0 V

Idrain

Vgate-to-source = 0.5 V

Vgate-to-source = 1 V
Vgate-to-source = 2 V (pinch-off)
Edrain-to-source
"Ohmic regions"

Ibase = 75 A

Icollector

Ibase = 40 A
Ibase = 20 A
Ibase = 5 A
Ecollector-to-emitter

A JFET transistor operated in the triode region tends to act very much like a plain resistor as
measured from drain to source. Like all simple resistances, its current/voltage graph is a straight
line. For this reason, the triode region (non-horizontal) portion of a JFETs characteristic curve
is sometimes referred to as the ohmic region. In this mode of operation where there isnt enough

264

CHAPTER 5. JUNCTION FIELD-EFFECT TRANSISTORS

drain-to-source voltage to bring drain current up to the regulated point, the drain current is directly
proportional to the drain-to-source voltage. In a carefully designed circuit, this phenomenon can be
used to an advantage. Operated in this region of the curve, the JFET acts like a voltage-controlled
resistance rather than a voltage-controlled current regulator, and the appropriate model for the
transistor is different:

D
G
S
N-channel JFET diode-rheostat model
(for saturation, or "ohmic," mode only!)
D

S
Here and here alone the rheostat (variable resistor) model of a transistor is accurate. It must
be remembered, however, that this model of the transistor holds true only for a narrow range of
its operation: when it is extremely saturated (far less voltage applied between drain and source
than what is needed to achieve full regulated current through the drain). The amount of resistance
(measured in ohms) between drain and source in this mode is controlled by how much reverse-bias
voltage is applied between gate and source. The less gate-to-source voltage, the less resistance
(steeper line on graph).
Because JFETs are voltage-controlled current regulators (at least when theyre allowed to operate
in their active), their inherent amplification factor cannot be expressed as a unitless ratio as with
BJTs. In other words, there is no ratio for a JFET. This is true for all voltage-controlled active
devices, including other types of field-effect transistors and even electron tubes. There is, however,
an expression of controlled (drain) current to controlling (gate-source) voltage, and it is called
transconductance. Its unit is Siemens, the same unit for conductance (formerly known as the mho).
Why this choice of units? Because the equation takes on the general form of current (output
signal) divided by voltage (input signal).

5.5.

THE COMMON-SOURCE AMPLIFIER PENDING

gfs =

265

ID
VGS

Where,
gfs = Transconductance in Siemens
ID = Change in drain current
VGS = Change in gate-source voltage
Unfortunately, the transconductance value for any JFET is not a stable quantity: it varies
significantly with the amount of gate-to-source control voltage applied to the transistor. As we
saw in the SPICE simulations, the drain current does not change proportionally with changes in
gate-source voltage. To calculate drain current for any given gate-source voltage, there is another
equation that may be used. It is obviously nonlinear upon inspection (note the power of 2), reflecting
the nonlinear behavior weve already experienced in simulation:

ID = IDSS ( 1 -

VGS
VGS(cutoff)

Where,
ID = Drain current
IDSS = Drain current with gate shorted to source
VGS = Gate-to-source voltage
VGS(cutoff) = Pinch-off gate-to-source voltage
REVIEW:
In their active modes, JFETs regulate drain current according to the amount of reverse-bias
voltage applied between gate and source, much like a BJT regulates collector current according
to base current. The mathematical ratio between drain current (output) and gate-to-source
voltage (input) is called transconductance, and it is measured in units of Siemens.
The relationship between gate-source (control) voltage and drain (controlled) current is nonlinear: as gate-source voltage is decreased, drain current increases exponentially. That is to
say, the transconductance of a JFET is not constant over its range of operation.
In their triode region, JFETs regulate drain-to-source resistance according to the amount of
reverse-bias voltage applied between gate and source. In other words, they act like voltagecontrolled resistances.

5.5

The common-source amplifier PENDING

*** PENDING ***


REVIEW:

266

CHAPTER 5. JUNCTION FIELD-EFFECT TRANSISTORS

5.6

The common-drain amplifier PENDING

*** PENDING ***


REVIEW:

5.7

The common-gate amplifier PENDING

*** PENDING ***


REVIEW:

5.8

Biasing techniques PENDING

*** PENDING ***


REVIEW:

5.9

Transistor ratings and packages PENDING

*** PENDING ***


REVIEW:

5.10.

5.10

JFET QUIRKS PENDING

JFET quirks PENDING

*** PENDING ***


REVIEW:

267

268

CHAPTER 5. JUNCTION FIELD-EFFECT TRANSISTORS

Chapter 6

INSULATED-GATE
FIELD-EFFECT TRANSISTORS
Contents
6.1
6.2
6.3
6.4
6.5
6.6
6.7
6.8
6.9
6.10
6.11
6.12

Introduction . . . . . . . . . . . . . . . . . . . .
Depletion-type IGFETs . . . . . . . . . . . . .
Enhancement-type IGFETs PENDING . .
Active-mode operation PENDING . . . . .
The common-source amplifier PENDING .
The common-drain amplifier PENDING . .
The common-gate amplifier PENDING . .
Biasing techniques PENDING . . . . . . . .
Transistor ratings and packages PENDING
IGFET quirks PENDING . . . . . . . . . . .
MESFETs PENDING . . . . . . . . . . . . .
IGBTs . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

269
270
279
279
280
280
280
280
280
281
281
281

*** INCOMPLETE ***

6.1

Introduction

As was stated in the last chapter, there is more than one type of field-effect transistor. The junction
field-effect transistor, or JFET, uses voltage applied across a reverse-biased PN junction to control
the width of that junctions depletion region, which then controls the conductivity of a semiconductor channel through which the controlled current moves. Another type of field-effect device the
insulated gate field-effect transistor, or IGFET exploits a similar principle of a depletion region
controlling conductivity through a semiconductor channel, but it differs primarily from the JFET
in that there is no direct connection between the gate lead and the semiconductor material itself.
Rather, the gate lead is insulated from the transistor body by a thin barrier, hence the term insulated
269

270

CHAPTER 6. INSULATED-GATE FIELD-EFFECT TRANSISTORS

gate. This insulating barrier acts like the dielectric layer of a capacitor, and allows gate-to-source
voltage to influence the depletion region electrostatically rather than by direct connection.
In addition to a choice of N-channel versus P-channel design, IGFETs come in two major types:
enhancement and depletion. The depletion type is more closely related to the JFET, so we will
begin our study of IGFETs with it.

6.2

Depletion-type IGFETs

Insulated gate field-effect transistors are unipolar devices just like JFETs: that is, the controlled
current does not have to cross a PN junction. There is a PN junction inside the transistor, but
its only purpose is to provide that nonconducting depletion region which is used to restrict current
through the channel.
Here is a diagram of an N-channel IGFET of the depletion type:

N-channel, D-type IGFET


drain
drain
substrate

gate
source

gate

N P

substrate

insulating
barrier
source

schematic symbol

physical diagram

Notice how the source and drain leads connect to either end of the N channel, and how the gate
lead attaches to a metal plate separated from the channel by a thin insulating barrier. That barrier
is sometimes made from silicon dioxide (the primary chemical compound found in sand), which
is a very good insulator. Due to this Metal (gate) - Oxide (barrier) - Semiconductor (channel)
construction, the IGFET is sometimes referred to as a MOSFET. There are other types of IGFET
construction, though, and so IGFET is the better descriptor for this general class of transistors.
Notice also how there are four connections to the IGFET. In practice, the substrate lead is directly
connected to the source lead to make the two electrically common. Usually, this connection is made
internally to the IGFET, eliminating the separate substrate connection, resulting in a three-terminal
device with a slightly different schematic symbol:

6.2.

DEPLETION-TYPE IGFETS

271

N-channel, D-type IGFET


drain
drain
gate

gate

N P

substrate

insulating
barrier

source

source
schematic symbol

physical diagram

With source and substrate common to each other, the N and P layers of the IGFET end up being
directly connected to each other through the outside wire. This connection prevents any voltage
from being impressed across the PN junction. As a result, a depletion region exists between the two
materials, but it can never be expanded or collapsed. JFET operation is based on the expansion of
the PN junctions depletion region, but here in the IGFET that cannot happen, so IGFET operation
must be based on a different effect.
Indeed it is, for when a controlling voltage is applied between gate and source, the conductivity of
the channel is changed as a result of the depletion region moving closer to or further away from the
gate. In other words, the channels effective width changes just as with the JFET, but this change
in channel width is due to depletion region displacement rather than depletion region expansion.
In an N-channel IGFET, a controlling voltage applied positive (+) to the gate and negative (-)
to the source has the effect of repelling the PN junctions depletion region, expanding the N-type
channel and increasing conductivity:

Rload
drain

gate
N
controlling
voltage

source

Channel expands for greater conductivity


Reversing the controlling voltages polarity has the opposite effect, attracting the depletion region
and narrowing the channel, consequently reducing channel conductivity:

272

CHAPTER 6. INSULATED-GATE FIELD-EFFECT TRANSISTORS

Rload
drain

gate

N
P

controlling
voltage

source

Channel narrows for less conductivity

The insulated gate allows for controlling voltages of any polarity without danger of forwardbiasing a junction, as was the concern with JFETs. This type of IGFET, although its called a
depletion-type, actually has the capability of having its channel either depleted (channel narrowed)
or enhanced (channel expanded). Input voltage polarity determines which way the channel will be
influenced.

Understanding which polarity has which effect is not as difficult as it may seem. The key is
to consider the type of semiconductor doping used in the channel (N-channel or P-channel?), then
relate that doping type to the side of the input voltage source connected to the channel by means of
the source lead. If the IGFET is an N-channel and the input voltage is connected so that the positive
(+) side is on the gate while the negative (-) side is on the source, the channel will be enhanced as
extra electrons build up on the channel side of the dielectric barrier. Think, negative (-) correlates
with N -type, thus enhancing the channel with the right type of charge carrier (electrons) and making
it more conductive. Conversely, if the input voltage is connected to an N-channel IGFET the other
way, so that negative (-) connects to the gate while positive (+) connects to the source, free electrons
will be robbed from the channel as the gate-channel capacitor charges, thus depleting the channel
of majority charge carriers and making it less conductive.

For P-channel IGFETs, the input voltage polarity and channel effects follow the same rule. That
is to say, it takes just the opposite polarity as an N-channel IGFET to either deplete or enhance:

6.2.

DEPLETION-TYPE IGFETS

273

Rload
drain

gate
P
controlling
voltage

source

Channel expands for greater conductivity

Rload
drain

gate

controlling
voltage

P
N

source

Channel narrows for less conductivity


Illustrating the proper biasing polarities with standard IGFET symbols:

274

CHAPTER 6. INSULATED-GATE FIELD-EFFECT TRANSISTORS

N-channel

P-channel

Enhanced
(more drain
current)

Depleted
(less drain
current)

When there is zero voltage applied between gate and source, the IGFET will conduct current
between source and drain, but not as much current as it would if it were enhanced by the proper
gate voltage. This places the depletion-type, or simply D-type, IGFET in a category of its own in the
transistor world. Bipolar junction transistors are normally-off devices: with no base current, they
block any current from going through the collector. Junction field-effect transistors are normally-on
devices: with zero applied gate-to-source voltage, they allow maximum drain current (actually, you
can coax a JFET into greater drain currents by applying a very small forward-bias voltage between
gate and source, but this should never be done in practice for risk of damaging its fragile PN
junction). D-type IGFETs, however, are normally half-on devices: with no gate-to-source voltage,
their conduction level is somewhere between cutoff and full saturation. Also, they will tolerate
applied gate-source voltages of any polarity, the PN junction being immune from damage due to the
insulating barrier and especially the direct connection between source and substrate preventing any
voltage differential across the junction.

Ironically, the conduction behavior of a D-type IGFET is strikingly similar to that of an electron
tube of the triode/tetrode/pentode variety. These devices were voltage-controlled current regulators
that likewise allowed current through them with zero controlling voltage applied. A controlling
voltage of one polarity (grid negative and cathode positive) would diminish conductivity through
the tube while a voltage of the other polarity (grid positive and cathode negative) would enhance
conductivity. I find it curious that one of the later transistor designs invented exhibits the same
basic properties of the very first active (electronic) device.

A few SPICE analyses will demonstrate the current-regulating behavior of D-type IGFETs. First,
a test with zero input voltage (gate shorted to source) and the power supply swept from 0 to 50
volts. The graph shows drain current:

6.2.

DEPLETION-TYPE IGFETS

275

Vammeter
1

2
0V
V1

Q1

0
0

n-channel igfet characteristic curve


m1 1 0 0 0 mod1
vammeter 2 1 dc 0
v1 2 0
.model mod1 nmos vto=-1
.dc v1 0 50 2
.plot dc i(vammeter)
.end

As expected for any transistor, the controlled current holds steady at a regulated value over a
wide range of power supply voltages. In this case, that regulated point is 10 A (1.000E-05). Now
lets see what happens when we apply a negative voltage to the gate (with reference to the source)
and sweep the power supply over the same range of 0 to 50 volts:

276

CHAPTER 6. INSULATED-GATE FIELD-EFFECT TRANSISTORS

Vammeter
1

2
0V
V1

Q1

0.5 V

n-channel igfet characteristic curve


m1 1 3 0 0 mod1
vin 0 3 dc 0.5
vammeter 2 1 dc 0
v1 2 0
.model mod1 nmos vto=-1
.dc v1 0 50 2
.plot dc i(vammeter)
.end

Not surprisingly, the drain current is now regulated at a lower value of 2.5 A (down from 10
A with zero input voltage). Now lets apply an input voltage of the other polarity, to enhance the
IGFET:

6.2.

DEPLETION-TYPE IGFETS

277

Vammeter
1

2
0V
V1

Q1

0.5 V

n-channel igfet characteristic curve


m1 1 3 0 0 mod1
vin 3 0 dc 0.5
vammeter 2 1 dc 0
v1 2 0
.model mod1 nmos vto=-1
.dc v1 0 50 2
.plot dc i(vammeter)
.end

With the transistor enhanced by the small controlling voltage, the drain current is now at an
increased value of 22.5 A (2.250E-05). It should be apparent from these three sets of voltage and
current figures that the relationship of drain current to gate-source voltage is nonlinear just as it
was with the JFET. With 1/2 volt of depleting voltage, the drain current is 2.5 A; with 0 volts
input the drain current goes up to 10 A; and with 1/2 volt of enhancing voltage, the current is
at 22.5 A. To obtain a better understanding of this nonlinearity, we can use SPICE to plot the

278

CHAPTER 6. INSULATED-GATE FIELD-EFFECT TRANSISTORS

drain current over a range of input voltage values, sweeping from a negative (depleting) figure to a
positive (enhancing) figure, maintaining the power supply voltage of V 1 at a constant value:

n-channel igfet
m1 1 3 0 0 mod1
vin 3 0
vammeter 2 1 dc 0
v1 2 0 dc 24
.model mod1 nmos vto=-1
.dc vin -1 1 0.1
.plot dc i(vammeter)
.end

Just as it was with JFETs, this inherent nonlinearity of the IGFET has the potential to cause
distortion in an amplifier circuit, as the input signal will not be reproduced with 100 percent accuracy
at the output. Also notice that a gate-source voltage of about 1 volt in the depleting direction is
able to pinch off the channel so that there is virtually no drain current. D-type IGFETs, like JFETs,
have a certain pinch-off voltage rating. This rating varies with the precise unique of the transistor,
and may not be the same as in our simulation here.
Plotting a set of characteristic curves for the IGFET, we see a pattern not unlike that of the
JFET:

6.3.

ENHANCEMENT-TYPE IGFETS PENDING

Vgate-to-source = +0.5 V

Idrain
Vgate-to-source = 0 V

Vgate-to-source = -0.5 V
Edrain-to-source
REVIEW:

6.3

Enhancement-type IGFETs PENDING

REVIEW:

6.4

Active-mode operation PENDING

REVIEW:

279

280

CHAPTER 6. INSULATED-GATE FIELD-EFFECT TRANSISTORS

6.5

The common-source amplifier PENDING

REVIEW:

6.6

The common-drain amplifier PENDING

REVIEW:

6.7

The common-gate amplifier PENDING

REVIEW:

6.8

Biasing techniques PENDING

REVIEW:

6.9

Transistor ratings and packages PENDING

REVIEW:

6.10.

6.10

IGFET QUIRKS PENDING

281

IGFET quirks PENDING

REVIEW:

6.11

MESFETs PENDING

REVIEW:

6.12

IGBTs

Because of their insulated gates, IGFETs of all types have extremely high current gain: there can be
no sustained gate current if there is no continuous gate circuit in which electrons may continually
flow. The only current we see through the gate terminal of an IGFET, then, is whatever transient
(brief surge) may be required to charge the gate-channel capacitance and displace the depletion
region as the transistor switches from an on state to an off state, or vice versa.
This high current gain would at first seem to place IGFET technology at a decided advantage
over bipolar transistors for the control of very large currents. If a bipolar junction transistor is used
to control a large collector current, there must be a substantial base current sourced or sunk by
some control circuitry, in accordance with the ratio. To give an example, in order for a power
BJT with a of 20 to conduct a collector current of 100 amps, there must be at least 5 amps of
base current, a substantial amount of current in itself for miniature discrete or integrated control
circuitry to handle:

Rload
100 A
Control
circuitry

5A

= 20
105 A

282

CHAPTER 6. INSULATED-GATE FIELD-EFFECT TRANSISTORS

It would be nice from the standpoint of control circuitry to have power transistors with high
current gain, so that far less current is needed for control of load current. Of course, we can use
Darlington pair transistors to increase the current gain, but this kind of arrangement still requires
far more controlling current than an equivalent power IGFET:

Rload
0.238 A
100 A

Control
circuitry

= 20
5A
105 A

Rload
100 A
Control
circuitry

0A
100 A

Unfortunately, though, IGFETs have problems of their own controlling high current: they typically exhibit greater drain-to-source voltage drop while saturated than the collector-to-emitter voltage drop of a saturated BJT. This greater voltage drop equates to higher power dissipation for the
same amount of load current, limiting the usefulness of IGFETs as high-power devices. Although
some specialized designs such as the so-called VMOS transistor have been designed to minimize this
inherent disadvantage, the bipolar junction transistor is still superior in its ability to switch high
currents.
An interesting solution to this dilemma leverages the best features of IGFETs with the best of
features of BJTs, in one device called an Insulated-Gate Bipolar Transistor, or IGBT. Also known
as an Bipolar-mode MOSFET, a Conductivity-Modulated Field-Effect Transistor (COMFET ), or
simply as an Insulated-Gate Transistor (IGT ), it is equivalent to a Darlington pair of IGFET and
BJT:

6.12.

IGBTS

283

Insulated-Gate Bipolar Transistor (IGBT)


(N-channel)
Schematic symbols
Collector

Equivalent circuit

Collector

Collector

Gate
Gate
Emitter

Emitter

Gate
Emitter

In essence, the IGFET controls the base current of a BJT, which handles the main load current
between collector and emitter. This way, there is extremely high current gain (since the insulated
gate of the IGFET draws practically no current from the control circuitry), but the collector-toemitter voltage drop during full conduction is as low as that of an ordinary BJT.
One disadvantage of the IGBT over a standard BJT is its slower turn-off time. For fast switching
and high current-handling capacity, its difficult to beat the bipolar junction transistor. Faster turnoff times for the IGBT may be achieved by certain changes in design, but only at the expense of a
higher saturated voltage drop between collector and emitter. However, the IGBT provides a good
alternative to IGFETs and BJTs for high-power control applications.
REVIEW:

284

CHAPTER 6. INSULATED-GATE FIELD-EFFECT TRANSISTORS

Chapter 7

THYRISTORS
Contents
7.1

Hysteresis

7.2

Gas discharge tubes

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285

7.3

The Shockley Diode . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 290

7.4

The DIAC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 296

7.5

The Silicon-Controlled Rectifier (SCR) . . . . . . . . . . . . . . . . . . 297

7.6

The TRIAC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308

7.7

Optothyristors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310

7.8

The Unijunction Transistor (UJT) PENDING . . . . . . . . . . . . 311

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286

7.9

The Silicon-Controlled Switch (SCS) . . . . . . . . . . . . . . . . . . . 311

7.10

Field-effect-controlled thyristors . . . . . . . . . . . . . . . . . . . . . . 313

*** INCOMPLETE ***

7.1

Hysteresis

Thyristors are a class of semiconductor components exhibiting hysteresis, that property whereby a
system fails to return to its original state after some cause of state change has been removed. A very
simple example of hysteresis is the mechanical action of a toggle switch: when the lever is pushed, it
flips to one of two extreme states (positions) and will remain there even after the source of motion is
removed (after you remove your hand from the switch lever). To illustrate the absence of hysteresis,
consider the action of a momentary pushbutton switch, which returns to its original state after
the button is no longer pressed: when the stimulus is removed (your hand), the system (switch)
immediately and fully returns to its prior state with no latching behavior.
Bipolar, junction field-effect, and insulated gate field-effect transistors are all non-hysteretic devices. That is, they do not inherently latch into a state after being stimulated by a voltage or
current signal. For any given input signal at any given time, a transistor will exhibit a predictable
285

286

CHAPTER 7. THYRISTORS

output response as defined by its characteristic curve. Thyristors, on the other hand, are semiconductor devices that tend to stay on once turned on, and tend to stay off once turned off. A
momentary event is able to flip these devices into either their on or off states where they will remain
that way on their own, even after the cause of the state change is taken away. As such, they are
useful only as on/off switching devices much like a toggle switch and cannot be used as analog
signal amplifiers.
Thyristors are constructed using the same technology as bipolar junction transistors, and in fact
may be analyzed as circuits comprised of transistor pairs. How then, can a hysteretic device (a
thyristor) be made from non-hysteretic devices (transistors)? The answer to this question is positive
feedback, also known as regenerative feedback. As you should recall, feedback is the condition where
a percentage of the output signal is fed back to the input of an amplifying device. Negative, or
degenerative, feedback results in a diminishing of voltage gain with increases in stability, linearity,
and bandwidth. Positive feedback, on the other hand, results in a kind of instability where the
amplifiers output tends to saturate. In the case of thyristors, this saturating tendency equates to
the device wanting to stay on once turned on, and off once turned off.
In this chapter we will explore several different kinds of thyristors, most of which stem from a
single, basic two-transistor core circuit. Before we do that, though, it would be beneficial to study
the technological predecessor to thyristors: gas discharge tubes.

7.2

Gas discharge tubes

If youve ever witnessed a lightning storm, youve seen electrical hysteresis in action (and probably
didnt realize what you were seeing). The action of strong wind and rain accumulates tremendous
static electric charges between cloud and earth, and between clouds as well. Electric charge imbalances manifest themselves as high voltages, and when the electrical resistance of air can no longer
hold these high voltages at bay, huge surges of current travel between opposing poles of electrical
charge which we call lightning.
The buildup of high voltages by wind and rain is a fairly continuous process, the rate of charge
accumulation increasing under the proper atmospheric conditions. However, lightning bolts are
anything but continuous: they exist as relatively brief surges rather than continuous discharges.
Why is this? Why dont we see soft, glowing lightning arcs instead of violently brief lightning bolts?
The answer lies in the nonlinear (and hysteretic) resistance of air.
Under ordinary conditions, air has an extremely high amount of resistance. It is so high, in
fact, that we typically treat its resistance as infinite and electrical conduction through the air as
negligible. The presence of water and/or dust in air lowers its resistance some, but it is still an
insulator for most practical purposes. When a sufficient amount of high voltage is applied across
a distance of air, though, its electrical properties change: electrons become stripped from their
normal positions around their respective atoms and are liberated to constitute a current. In this
state, air is considered to be ionized and is referred to as a plasma rather than a normal gas. This
usage of the word plasma is not to be confused with the medical term (meaning the fluid portion
of blood), but is a fourth state of matter, the other three being solid, liquid, and vapor (gas). Plasma
is a relatively good conductor of electricity, its specific resistance being much lower than that of the
same substance in its gaseous state.
As an electric current moves through the plasma, there is energy dissipated in the plasma in
the form of heat, just as current through a solid resistor dissipates energy in the form of heat. In

7.2.

GAS DISCHARGE TUBES

287

the case of lightning, the temperatures involved are extremely high. High temperatures are also
sufficient to convert gaseous air into a plasma or maintain plasma in that state without the presence
of high voltage. As the voltage between cloud and earth, or between cloud and cloud, decreases as
the charge imbalance is neutralized by the current of the lightning bolt, the heat dissipated by the
bolt maintains the air path in a plasma state, keeping its resistance low. The lightning bolt remains
a plasma until the voltage decreases to too low a level to sustain enough current to dissipate enough
heat. Finally, the air returns to a normal, gaseous state and stops conducting current, thus allowing
voltage to build up once more.
Note how throughout this cycle, the air exhibits hysteresis. When not conducting electricity, it
tends to remain an insulator until voltage builds up past a critical threshold point. Then, once it
changes state and becomes a plasma, it tends to remain a conductor until voltage falls below a lower
critical threshold point. Once turned on it tends to stay on, and once turned off it tends to
stay off. This hysteresis, combined with a steady buildup of voltage due to the electrostatic effects
of wind and rain, explains the action of lightning as brief bursts.
In electronic terms, what we have here in the action of lightning is a simple relaxation oscillator.
Oscillators are electronic circuits that produce an oscillating (AC) voltage from a steady supply of
DC power. A relaxation oscillator is one that works on the principle of a charging capacitor that
is suddenly discharged every time its voltage reaches a critical threshold value. One of the simplest
relaxation oscillators in existence is comprised of three components (not counting the DC power
supply): a resistor, capacitor, and neon lamp:

Simple relaxation oscillator


R

Neon lamp

Neon lamps are nothing more than two metal electrodes inside a sealed glass bulb, separated by
the neon gas inside. At room temperatures and with no applied voltage, the lamp has nearly infinite
resistance. However, once a certain threshold voltage is exceeded (this voltage depends on the gas
pressure and geometry of the lamp), the neon gas will become ionized (turned into a plasma) and
its resistance dramatically reduced. In effect, the neon lamp exhibits the same characteristics as air
in a lightning storm, complete with the emission of light as a result of the discharge, albeit on a
much smaller scale.
The capacitor in the relaxation oscillator circuit shown above charges at an inverse exponential
rate determined by the size of the resistor. When its voltage reaches the threshold voltage of the
lamp, the lamp suddenly turns on and quickly discharges the capacitor to a low voltage value.
Once discharged, the lamp turns off and allows the capacitor to build up a charge once more. The
result is a series of brief flashes of light from the lamp, the rate of which dictated by battery voltage,

288

CHAPTER 7. THYRISTORS

resistor resistance, capacitor capacitance, and lamp threshold voltage.


While gas-discharge lamps are more commonly used as sources of illumination, their hysteretic
properties were leveraged in slightly more sophisticated variants known as thyratron tubes. Essentially a gas-filled triode tube (a triode being a three-element vacuum electron tube performing
much a similar function to the N-channel, D-type IGFET), the thyratron tube could be turned
on with a small control voltage applied between grid and cathode, and turned off by reducing the
plate-to-cathode voltage.
(Simple) Thyratron control circuit
+
Rload
Thyratron
Tube

high voltage
AC source

control
voltage

In essence, thyratron tubes were controlled versions of neon lamps built specifically for switching
current to a load. The dot inside the circle of the schematic symbol indicates a gas fill, as opposed
to the hard vacuum normally seen in other electron tube designs. In the circuit shown above, the
thyratron tube allows current through the load in one direction (note the polarity across the load
resistor) when triggered by the small DC control voltage connected between grid and cathode. Note
that the loads power source is AC, which provides a clue as to how the thyratron turns off after
its been triggered on: since AC voltage periodically passes through a condition of 0 volts between
half-cycles, the current through an AC-powered load must also periodically halt. This brief pause
of current between half-cycles gives the tubes gas time to cool, letting it return to its normal off
state. Conduction may resume only if there is enough voltage applied by the AC power source (some
other time in the waves cycle) and if the DC control voltage allows it.
An oscilloscope display of load voltage in such a circuit would look something like this:

Threshold voltage

AC supply voltage
Load voltage

7.2.

GAS DISCHARGE TUBES

289

As the AC supply voltage climbs from zero volts to its first peak, the load voltage remains at
zero (no load current) until the threshold voltage is reached. At that point, the tube switches on
and begins to conduct, the load voltage now following the AC voltage through the rest of the half
cycle. Notice how there is load voltage (and thus load current) even when the AC voltage waveform
has dropped below the threshold value of the tube. This is hysteresis at work: the tube stays in its
conductive mode past the point where it first turned on, continuing to conduct until there the supply
voltage drops off to almost zero volts. Because thyratron tubes are one-way (diode) devices, there
is no voltage across the load through the negative half-cycle of AC. In practical thyratron circuits,
multiple tubes arranged in some form of full-wave rectifier circuit to facilitate full-wave DC power
to the load.
Although Im not sure if this was ever done, someone could have applied the thyratron tube
to a relaxation oscillator circuit and control the frequency with a small DC voltage between grid
and cathode, making a crude voltage-controlled oscillator, otherwise known as a VCO. Relaxation
oscillators tend to have poor frequency control, not to mention a very non-sinusoidal output, and
so they exist mostly as demonstration circuits (as is the case here) or in applications where precise
frequency control isnt important. Consequently, this use of a thyratron tube would not have been
a very practical one.

C
Controlling
voltage
I speak of thyratron tubes in the past tense for good reason: modern semiconductor components
have obsoleted thyratron tube technology for all but a few very special applications. It is no coincidence that the word thyristor bears so much similarity to the word thyratron, for this class of
semiconductor components does much the same thing: use hyster etically switch current on and off.
It is these modern devices that we now turn our attention to.
REVIEW:
Electrical hysteresis, the tendency for a component to remain on (conducting) after it begins
to conduct and to remain off (nonconducting) after it ceases to conduct, helps to explain
why lightning bolts exist as momentary surges of current rather than continuous discharges
through the air.
Simple gas-discharge tubes such as neon lamps exhibit electrical hysteresis.
More advanced gas-discharge tubes have been made with control elements so that their turnon voltage could be adjusted by an external signal. The most common of these tubes was
called the thyratron.

290

CHAPTER 7. THYRISTORS
Simple oscillator circuits called relaxation oscillators may be created with nothing more than
a resistor-capacitor charging network and a hysteretic device connected across the capacitor.

7.3

The Shockley Diode

Our exploration of thyristors begins with a device called the four-layer diode, also known as a PNPN
diode, or a Shockley diode after its inventor, William Shockley. This is not to be confused with a
Schottky diode, that two-layer metal-semiconductor device known for its high switching speed. A
crude illustration of the Shockley diode, often seen in textbooks, is a four-layer sandwich of P-N-P-N
semiconductor material:

Shockley, or 4-layer,
diode

Anode

N
P
N

Cathode

Unfortunately, this simple illustration does nothing to enlighten the viewer on how it works or
why. Consider an alternative rendering of the devices construction:

Shockley, or 4-layer,
diode

Anode

P
N

Cathode

Shown like this, it appears to be a set of interconnected bipolar transistors, one PNP and the
other NPN. Drawn using standard schematic symbols, and respecting the layer doping concentrations
not shown in the last image, the Shockley diode looks like this:

7.3.

THE SHOCKLEY DIODE


Anode

291

Anode
P
N
P

N
P
N

Cathode
Physical diagram

Cathode
Equivalent schematic

Schematic symbol

Lets connect one of these devices to a source of variable voltage and see what happens:

Shockley diode
equivalent circuit

With no voltage applied, of course there will be no current. As voltage is initially increased,
there will still be no current because neither transistor is able to turn on: both will be in cutoff
mode. To understand why this is, consider what it takes to turn a bipolar junction transistor on:
current through the base-emitter junction. As you can see in the diagram, base current through
the lower transistor is controlled by the upper transistor, and the base current through the upper
transistor is controlled by the lower transistor. In other words, neither transistor can turn on until
the other transistor turns on. What we have here, in vernacular terms, is known as a Catch-22.
So how can a Shockley diode ever conduct current, if its constituent transistors stubbornly
maintain themselves in a state of cutoff? The answer lies in the behavior of real transistors as
opposed to ideal transistors. An ideal bipolar transistor will never conduct collector current if there
is no base current, no matter how much or little voltage we apply between collector and emitter.
Real transistors, on the other hand, have definite limits to how much collector-emitter voltage they
can withstand before they break down and conduct. If two real transistors are connected together
in this fashion to form a Shockley diode, they will be able to conduct if there is sufficient voltage
applied by the battery between anode and cathode to cause one of them to break down. Once one
transistor breaks down and begins to conduct, it will allow base current through the other transistor,
causing it to turn on in a normal fashion, which then allows base current through the first transistor.
The end result is that both transistors will be saturated, now keeping each other turned on instead
of off.
So, we can force a Shockley diode to turn on by applying sufficient voltage between anode and
cathode. As we have seen, this will inevitably cause one of the transistors to turn on, which then turns
the other transistor on, ultimately latching both transistors on where they will tend to remain.
But how do we now get the two transistors to turn off again? Even if the applied voltage is reduced

292

CHAPTER 7. THYRISTORS

to a point well below what it took to get the Shockley diode conducting, it will remain conducting
because both transistors now have base current to maintain regular, controlled conduction. The
answer to this is to reduce the applied voltage to a much lower point where there is too little current
to maintain transistor bias, at which point one of the transistors will cutoff, which then halts base
current through the other transistor, sealing both transistors in the off state as they were before
any voltage was applied at all.
If we graph this sequence of events and plot the results on an I/V graph, the hysteresis is very
evident. First, we will observe the circuit as the DC voltage source (battery) is set to zero voltage:

Circuit
current

Applied voltage

Zero applied voltage; zero current


Next, we will steadily increase the DC voltage. Current through the circuit is at or nearly at
zero, as the breakdown limit has not been reached for either transistor:

Circuit
current

Applied voltage

Some voltage applied, still no appreciable current


When the voltage breakdown limit of one transistor is reached, it will begin to conduct collector
current even though no base current has gone through it yet. Normally, this sort of treatment
would destroy a bipolar junction transistor, but the PNP junctions comprising a Shockley diode are
engineered to take this kind of abuse, similar to the way a Zener diode is built to handle reverse
breakdown without sustaining damage. For the sake of illustration Ill assume the lower transistor
breaks down first, sending current through the base of the upper transistor:

7.3.

THE SHOCKLEY DIODE

293

Circuit
current

Applied voltage
More voltage applied; lower transistor breaks down
As the upper transistor receives base current, it turns on as expected. This action allows the
lower transistor to conduct normally, the two transistors sealing themselves in the on state. Full
current is very quickly seen in the circuit:

Circuit
current

Applied voltage
Transistors now fully conducting
The positive feedback mentioned earlier in this chapter is clearly evident here. When one transistor breaks down, it allows current through the device structure. This current may be viewed as
the output signal of the device. Once an output current is established, it works to hold both
transistors in saturation, thus ensuring the continuation of a substantial output current. In other
words, an output current feeds back positively to the input (transistor base current) to keep both
transistors in the on state, thus reinforcing (or regenerating) itself.
With both transistors maintained in a state of saturation with the presence of ample base current,
they will continue to conduct even if the applied voltage is greatly reduced from the breakdown level.
The effect of positive feedback is to keep both transistors in a state of saturation despite the loss
of input stimulus (the original, high voltage needed to break down one transistor and cause a base
current through the other transistor):

294

CHAPTER 7. THYRISTORS

Circuit
current

Applied voltage
Current maintained even when voltage is reduced
If the DC voltage source is turned down too far, though, the circuit will eventually reach a point
where there isnt enough current to sustain both transistors in saturation. As one transistor passes
less and less collector current, it reduces the base current for the other transistor, thus reducing base
current for the first transistor. The vicious cycle continues rapidly until both transistors fall into
cutoff:

Circuit
current

Applied voltage
If the voltage drops too low, both transistors shut off
Here, positive feedback is again at work: the fact that the cause/effect cycle between both transistors is vicious (a decrease in current through one works to decrease current through the other,
further decreasing current through the first transistor) indicates a positive relationship between
output (controlled current) and input (controlling current through the transistors bases).
The resulting curve on the graph is classicly hysteretic: as the input signal (voltage) is increased
and decreased, the output (current) does not follow the same path going down as it did going up:

Circuit
current

Applied voltage

7.3.

THE SHOCKLEY DIODE

295

Put in simple terms, the Shockley diode tends to stay on once its turned on, and stay off once
its turned off. There is no in-between or active mode in its operation: it is a purely on or off
device, as are all thyristors.
There are a few special terms applied to Shockley diodes and all other thyristor devices built
upon the Shockley diode foundation. First is the term used to describe its on state: latched. The
word latch is reminiscent of a door lock mechanism, which tends to keep the door closed once it
has been pushed shut. The term firing refers to the initiation of a latched state. In order to get a
Shockley diode to latch, the applied voltage must be increased until breakover is attained. Despite
the fact that this action is best described in terms of transistor breakdown, the term breakover
is used instead because the end result is a pair of transistors in mutual saturation rather than
destruction as would be the case with a normal transistor. A latched Shockley diode is re-set back
into its nonconducting state by reducing current through it until low-current dropout occurs.
It should be noted that Shockley diodes may be fired in a way other than breakover: excessive
voltage rise, or dv/dt. This is when the applied voltage across the diode increases at a high rate of
change. This is able to cause latching (turning on) of the diode due to inherent junction capacitances
within the transistors. Capacitors, as you may recall, oppose changes in voltage by drawing or
supplying current. If the applied voltage across a Shockley diode rises at too fast a rate, those
tiny capacitances will draw enough current during that time to activate the transistor pair, turning
them both on. Usually, this form of latching is undesirable, and can be minimized by filtering highfrequency (fast voltage rises) from the diode with series inductors and/or parallel resistor-capacitor
networks called snubbers:

Series inductor

Shockley
diode

RC "snubber"

Both the series inductor and the parallel


resistor-capacitor "snubber" circuit help
minimize the Shockley diodes exposure
to excessively rising voltages.
The voltage rise limit of a Shockley diode is referred to as the critical rate of voltage rise.
Manufacturers usually provide this specification for the devices they sell.
REVIEW:
Shockley diodes are four-layer PNPN semiconductor devices. They behave as a pair of interconnected PNP and NPN transistors.

296

CHAPTER 7. THYRISTORS
Like all thyristors, Shockley diodes tend to stay on once theyve been turned on (latched ), and
stay off once theyve been turned off.
There are two ways to latch a Shockley diode: exceed the anode-to-cathode breakover voltage,
or exceed the anode-to-cathode critical rate of voltage rise.
There is only one way to cause a Shockley diode to stop conducting, and that is to reduce the
current going through it to a level below its low-current dropout threshold.

7.4

The DIAC

Like all diodes, Shockley diodes are unidirectional devices; that is, they only conduct current in one
direction. If bidirectional (AC) operation is desired, two Shockley diodes may be joined in parallel
facing different directions to form a new kind of thyristor, the DIAC :

DIAC equivalent circuit

DIAC schematic symbol

A DIAC operated with a DC voltage across it behaves exactly the same as a Shockley diode.
With AC, however, the behavior is different from what one might expect. Because alternating
current repeatedly reverses direction, DIACs will not stay latched longer than one-half cycle. If a
DIAC becomes latched, it will continue to conduct current only as long as there is voltage available
to push enough current in that direction. When the AC polarity reverses, as it must twice per
cycle, the DIAC will drop out due to insufficient current, necessitating another breakover before it
conducts again. The result is a current waveform that looks like this:

Breakover voltage

AC supply voltage

Breakover voltage

DIAC current
DIACs are almost never used alone, but in conjunction with other thyristor devices.

7.5.

THE SILICON-CONTROLLED RECTIFIER (SCR)

7.5

297

The Silicon-Controlled Rectifier (SCR)

Shockley diodes are curious devices, but rather limited in application. Their usefulness may be
expanded, however, by equipping them with another means of latching. In doing so, they become
true amplifying devices (if only in an on/off mode), and we refer to them as silicon-controlled
rectifiers, or SCRs.
The progression from Shockley diode to SCR is achieved with one small addition, actually nothing
more than a third wire connection to the existing PNPN structure:
The Silicon-Controlled Rectifier (SCR)

Anode

Anode
P
N

Gate

Anode
N
P
N

Cathode
Physical diagram

Gate

Gate
Cathode

Cathode
Equivalent schematic

Schematic symbol

If an SCRs gate is left floating (disconnected), it behaves exactly as a Shockley diode. It may
be latched by breakover voltage or by exceeding the critical rate of voltage rise between anode and
cathode, just as with the Shockley diode. Dropout is accomplished by reducing current until one or
both internal transistors fall into cutoff mode, also like the Shockley diode. However, because the
gate terminal connects directly to the base of the lower transistor, it may be used as an alternative
means to latch the SCR. By applying a small voltage between gate and cathode, the lower transistor
will be forced on by the resulting base current, which will cause the upper transistor to conduct,
which then supplies the lower transistors base with current so that it no longer needs to be activated
by a gate voltage. The necessary gate current to initiate latch-up, of course, will be much lower
than the current through the SCR from cathode to anode, so the SCR does achieve a measure of
amplification.
This method of securing SCR conduction is called triggering, and it is by far the most common
way that SCRs are latched in actual practice. In fact, SCRs are usually chosen so that their breakover
voltage is far beyond the greatest voltage expected to be experienced from the power source, so that
it can be turned on only by an intentional voltage pulse applied to the gate.
It should be mentioned that SCRs may sometimes be turned off by directly shorting their gate
and cathode terminals together, or by reverse-triggering the gate with a negative voltage (in
reference to the cathode), so that the lower transistor is forced into cutoff. I say this is sometimes
possible because it involves shunting all of the upper transistors collector current past the lower
transistors base. This current may be substantial, making triggered shut-off of an SCR difficult at
best. A variation of the SCR, called a Gate-Turn-Off thyristor, or GTO, makes this task easier.
But even with a GTO, the gate current required to turn it off may be as much as 20% of the anode
(load) current! The schematic symbol for a GTO is shown in the following illustration:

298

CHAPTER 7. THYRISTORS

Gate Turn-Off thyristor (GTO)


Anode
Gate
Cathode
SCRs and GTOs share the same equivalent schematics (two transistors connected in a positivefeedback fashion), the only differences being details of construction designed to grant the NPN
transistor a greater than the PNP. This allows a smaller gate current (forward or reverse) to exert
a greater degree of control over conduction from cathode to anode, with the PNP transistors latched
state being more dependent upon the NPNs than vice versa. The Gate-Turn-Off thyristor is also
known by the name of Gate-Controlled Switch, or GCS.
A rudimentary test of SCR function, or at least terminal identification, may be performed with
an ohmmeter. Because the internal connection between gate and cathode is a single PN junction, a
meter should indicate continuity between these terminals with the red test lead on the gate and the
black test lead on the cathode like this:

gate

cathode

A
OFF

COM

All other continuity measurements performed on an SCR will show open (OL on some digital
multimeter displays). It must be understood that this test is very crude and does not constitute a
comprehensive assessment of the SCR. It is possible for an SCR to give good ohmmeter indications
and still be defective. Ultimately, the only way to test an SCR is to subject it to a load current.
If you are using a multimeter with a diode check function, the gate-to-cathode junction voltage
indication you get may or may not correspond to whats expected of a silicon PN junction (approximately 0.7 volts). In some cases, you will read a much lower junction voltage: mere hundredths
of a volt. This is due to an internal resistor connected between the gate and cathode incorporated
within some SCRs. This resistor is added to make the SCR less susceptible to false triggering by
spurious voltage spikes, from circuit noise or from static electric discharge. In other words, having a resistor connected across the gate-cathode junction requires that a strong triggering signal
(substantial current) be applied to latch the SCR. This feature is often found in larger SCRs, not

7.5.

THE SILICON-CONTROLLED RECTIFIER (SCR)

299

on small SCRs. Bear in mind that an SCR with an internal resistor connected between gate and
cathode will indicate continuity in both directions between those two terminals:

Anode

Gate

Gate-to-Cathode
resistor

Cathode

Normal SCRs, lacking this internal resistor, are sometimes referred to as sensitive gate SCRs
due to their ability to be triggered by the slightest positive gate signal.
The test circuit for an SCR is both practical as a diagnostic tool for checking suspected SCRs and
also an excellent aid to understanding basic SCR operation. A DC voltage source is used for powering
the circuit, and two pushbutton switches are used to latch and unlatch the SCR, respectively:

SCR testing circuit


off

on

SCR under
test

Actuating the normally-open on pushbutton switch connects the gate to the anode, allowing
current from the negative terminal of the battery, through the cathode-gate PN junction, through
the switch, through the load resistor, and back to the battery. This gate current should force the
SCR to latch on, allowing current to go directly from cathode to anode without further triggering
through the gate. When the on pushbutton is released, the load should remain energized.
Pushing the normally-closed off pushbutton switch breaks the circuit, forcing current through
the SCR to halt, thus forcing it to turn off (low-current dropout).
If the SCR fails to latch, the problem may be with the load and not the SCR. There is a certain
minimum amount of load current required to hold the SCR latched in the on state. This minimum
current level is called the holding current. A load with too great a resistance value may not draw
enough current to keep an SCR latched when gate current ceases, thus giving the false impression
of a bad (unlatchable) SCR in the test circuit. Holding current values for different SCRs should

300

CHAPTER 7. THYRISTORS

be available from the manufacturers. Typical holding current values range from 1 milliamp to 50
milliamps or more for larger units.
For the test to be fully comprehensive, more than the triggering action needs to be tested. The
forward breakover voltage limit of the SCR could be tested by increasing the DC voltage supply
(with no pushbuttons actuated) until the SCR latches all on its own. Beware that a breakover test
may require very high voltage: many power SCRs have breakover voltage ratings of 600 volts or
more! Also, if a pulse voltage generator is available, the critical rate of voltage rise for the SCR
could be tested in the same way: subject it to pulsing supply voltages of different V/time rates with
no pushbutton switches actuated and see when it latches.
In this simple form, the SCR test circuit could suffice as a start/stop control circuit for a DC
motor, lamp, or other practical load:

DC motor start/stop
control circuit
Motor

off

on

SCR under
test

Another practical use for the SCR in a DC circuit is as a crowbar device for overvoltage protection.
A crowbar circuit consists of an SCR placed in parallel with the output of a DC power supply,
for the purpose of placing a direct short-circuit on the output of that supply to prevent excessive
voltage from reaching the load. Damage to the SCR and power supply is prevented by the judicious
placement of a fuse or substantial series resistance ahead of the SCR to limit short-circuit current:

Crowbar as used in an AC-DC power supply

Transformer
Filter

AC
power
source

Fuse

Load

Rectifier
Crowbar
(triggering circuit
omitted for simplicity)

Some device or circuit sensing the output voltage will be connected to the gate of the SCR, so
that when an overvoltage condition occurs, voltage will be applied between the gate and cathode,
triggering the SCR and forcing the fuse to blow. The effect will be approximately the same as

7.5.

THE SILICON-CONTROLLED RECTIFIER (SCR)

301

dropping a solid steel crowbar directly across the output terminals of the power supply, hence the
name of the circuit.
Most applications of the SCR are for AC power control, despite the fact that SCRs are inherently
DC (unidirectional) devices. If bidirectional circuit current is required, multiple SCRs may be used,
with one or more facing each direction to handle current through both half-cycles of the AC wave.
The primary reason SCRs are used at all for AC power control applications is the unique response
of a thyristor to an alternating current. As we saw in the case of the thyratron tube (the electron
tube version of the SCR) and the DIAC, a hysteretic device triggered on during a portion of an
AC half-cycle will latch and remain on throughout the remainder of the half-cycle until the AC
current decreases to zero, as it must to begin the next half-cycle. Just prior to the zero-crossover
point of the current waveform, the thyristor will turn off due to insufficient current (this behavior
is also known as natural commutation) and must be fired again during the next cycle. The result is
a circuit current equivalent to a chopped up sine wave. For review, here is the graph of a DIACs
response to an AC voltage whose peak exceeds the breakover voltage of the DIAC:

Breakover voltage

Breakover voltage

AC supply voltage
DIAC current

With the DIAC, that breakover voltage limit was a fixed quantity. With the SCR, we have
control over exactly when the device becomes latched by triggering the gate at any point in time
along the waveform. By connecting a suitable control circuit to the gate of an SCR, we can chop
the sine wave at any point to allow for time-proportioned power control to a load.
Take the following circuit as an example. Here, an SCR is positioned in a circuit to control power
to a load from an AC source:

Load
AC
source

SCR

Being a unidirectional (one-way) device, at most we can only deliver half-wave power to the
load, in the half-cycle of AC where the supply voltage polarity is positive on the top and negative on
the bottom. However, for demonstrating the basic concept of time-proportional control, this simple
circuit is better than one controlling full-wave power (which would require two SCRs).
With no triggering to the gate, and the AC source voltage well below the SCRs breakover voltage
rating, the SCR will never turn on. Connecting the SCR gate to the anode through a normal
rectifying diode (to prevent reverse current through the gate in the event of the SCR containing

302

CHAPTER 7. THYRISTORS

a built-in gate-cathode resistor), will allow the SCR to be triggered almost immediately at the
beginning of every positive half-cycle:

Gate connected directly to anode through a diode;


nearly complete half-wave current through load
Load
AC
source

Load current

AC source voltage
We can delay the triggering of the SCR, however, by inserting some resistance into the gate
circuit, thus increasing the amount of voltage drop required before there is enough gate current to
trigger the SCR. In other words, if we make it harder for electrons to flow through the gate by
adding a resistance, the AC voltage will have to reach a higher point in its cycle before there will
be enough gate current to turn the SCR on. The result looks like this:

Resistance inserted in gate circuit;


less than half-wave current through load
Load
AC
source

Load current

AC source voltage
With the half-sine wave chopped up to a greater degree by delayed triggering of the SCR, the
load receives less average power (power is delivered for less time throughout a cycle). By making
the series gate resistor variable, we can make adjustments to the time-proportioned power:

7.5.

THE SILICON-CONTROLLED RECTIFIER (SCR)

303

Load
AC
source
trigger
threshold

Increasing the resistance raises the threshold level,


causing less power to be delivered to the load.
Decreasing the resistance lowers the threshold level,
causing more power to be delivered to the load.
Unfortunately, this control scheme has a significant limitation. In using the AC source waveform
for our SCR triggering signal, we limit control to the first half of the waveforms half-cycle. In other
words, there is no way for us to wait until after the waves peak to trigger the SCR. This means we
can turn down the power only to the point where the SCR turns on at the very peak of the wave:

Circuit at minimum power setting


Load
AC
source
trigger
threshold

Raising the trigger threshold any more will cause the circuit to not trigger at all, since not even
the peak of the AC power voltage will be enough to trigger the SCR. The result will be no power to
the load.
An ingenious solution to this control dilemma is found in the addition of a phase-shifting capacitor
to the circuit:

304

CHAPTER 7. THYRISTORS

Load
AC
source

Capacitor voltage
The smaller waveform shown on the graph is voltage across the capacitor. For the sake of
illustrating the phase shift, Im assuming a condition of maximum control resistance where the SCR
is not triggering at all and there is no load current, save for what little current goes through the
control resistor and capacitor. This capacitor voltage will be phase-shifted anywhere from 0 o to
90o lagging behind the power source AC waveform. When this phase-shifted voltage reaches a high
enough level, the SCR will trigger.
Assuming there is periodically enough voltage across the capacitor to trigger the SCR, the resulting load current waveform will look something like this:

Load
AC
source
load
current

trigger
threshold

Capacitor voltage
Because the capacitor waveform is still rising after the main AC power waveform has reached its
peak, it becomes possible to trigger the SCR at a threshold level beyond that peak, thus chopping
the load current wave further than it was possible with the simpler circuit. In reality, the capacitor
voltage waveform is a bit more complex that what is shown here, its sinusoidal shape distorted every
time the SCR latches on. However, what Im trying to illustrate here is the delayed triggering action
gained with the phase-shifting RC network, and so a simplified, undistorted waveform serves the
purpose well.
SCRs may also be triggered, or fired, by more complex circuits. While the circuit previously
shown is sufficient for a simple application like a lamp control, large industrial motor controls often

7.5.

THE SILICON-CONTROLLED RECTIFIER (SCR)

305

rely on more sophisticated triggering methods. Sometimes, pulse transformers are used to couple
a triggering circuit to the gate and cathode of an SCR to provide electrical isolation between the
triggering and power circuits:

...
...
to triggering
circuit
...

pulse
transformer

SCR
to power
circuit
...

When multiple SCRs are used to control power, their cathodes are often not electrically common,
making it difficult to connect a single triggering circuit to all SCRs equally. An example of this is
the controlled bridge rectifier shown here:

Controlled bridge rectifier

SCR1

SCR2

SCR4

SCR3

Load

In any bridge rectifier circuit, the rectifying diodes (or in this case, the rectifying SCRs) must
conduct in opposite pairs. SCR1 and SCR3 must be fired simultaneously, and likewise SCR2 and
SCR4 must be fired together as a pair. As you will notice, though, these pairs of SCRs do not share
the same cathode connections, meaning that it would not work to simply parallel their respective
gate connections and connect a single voltage source to trigger both:

306

CHAPTER 7. THYRISTORS

This strategy will not work for


triggering SCR2 and SCR4
together as a pair!

SCR1

SCR2

triggering
voltage
(pulse voltage
source)

SCR4

SCR3

Load

Although the triggering voltage source shown will trigger SCR4 , it will not trigger SCR2 properly
because the two thyristors do not share a common cathode connection to reference that triggering
voltage. Pulse transformers connecting the two thyristor gates to a common triggering voltage source
will work, however:

SCR1

SCR2

SCR4

SCR3

pulse
voltage
source

Load

Bear in mind that this circuit only shows the gate connections for two out of the four SCRs.
Pulse transformers and triggering sources for SCR1 and SCR3 , as well as the details of the pulse
sources themselves, have been omitted for the sake of simplicity.
Controlled bridge rectifiers are not limited to single-phase designs. In most industrial control
systems, AC power is available in three-phase form for maximum efficiency, and solid-state control
circuits are built to take advantage of that. A three-phase controlled rectifier circuit built with
SCRs, without pulse transformers or triggering circuitry shown, would look like this:

7.5.

THE SILICON-CONTROLLED RECTIFIER (SCR)

307

3-phase source
Controlled
rectifier

+
Load
-

REVIEW:
A Silicon-Controlled Rectifier, or SCR, is essentially a Shockley diode with an extra terminal
added. This extra terminal is called the gate, and it is used to trigger the device into conduction
(latch it) by the application of a small voltage.
To trigger, or fire, an SCR, voltage must be applied between the gate and cathode, positive to
the gate and negative to the cathode. When testing an SCR, a momentary connection between
the gate and anode is sufficient in polarity, intensity, and duration to trigger it.
SCRs may be fired by intentional triggering of the gate terminal, excessive voltage (breakdown)
between anode and cathode, or excessive rate of voltage rise between anode and cathode.
SCRs may be turned off by anode current falling below the holding current value (low-current
dropout), or by reverse-firing the gate (applying a negative voltage to the gate). Reversefiring is only sometimes effective, and always involves high gate current.
A variant of the SCR, called a Gate-Turn-Off thyristor (GTO), is specifically designed to be
turned off by means of reverse triggering. Even then, reverse triggering requires fairly high
current: typically 20% of the anode current.
SCR terminals may be identified by a continuity meter: the only two terminals showing any
continuity between them at all should be the gate and cathode. Gate and cathode terminals
connect to a PN junction inside the SCR, so a continuity meter should obtain a diode-like
reading between these two terminals with the red (+) lead on the gate and the black (-) lead
on the cathode. Beware, though, that some large SCRs have an internal resistor connected
between gate and cathode, which will affect any continuity readings taken by a meter.
SCRs are true rectifiers: they only allow current through them in one direction. This means
they cannot be used alone for full-wave AC power control.
If the diodes in a rectifier circuit are replaced by SCRs, you have the makings of a controlled
rectifier circuit, whereby DC power to a load may be time-proportioned by triggering the SCRs
at different points along the AC power waveform.

308

CHAPTER 7. THYRISTORS

7.6

The TRIAC

SCRs are unidirectional (one-way) current devices, making them useful for controlling DC only.
If two SCRs are joined in back-to-back parallel fashion just like two Shockley diodes were joined
together to form a DIAC, we have a new device known as the TRIAC :

Main Terminal 2
(MT2)

Gate
Main Terminal 1
(MT1)
TRIAC equivalent circuit

Main Terminal 2
(MT2)
Gate
Main Terminal 1
(MT1)

TRIAC schematic symbol

Because individual SCRs are more flexible to use in advanced control systems, they are more
commonly seen in circuits like motor drives, while TRIACs are usually seen in simple, low-power
applications like household dimmer switches. A simple lamp dimmer circuit is shown here, complete
with the phase-shifting resistor-capacitor network necessary for after-peak firing.

Lamp

AC
source

TRIACs are notorious for not firing symmetrically. This means they usually wont trigger at
the exact same gate voltage level for one polarity as for the other. Generally speaking, this is
undesirable, because unsymmetrical firing results in a current waveform with a greater variety of
harmonic frequencies. Waveforms that are symmetrical above and below their average centerlines are
comprised of only odd-numbered harmonics. Unsymmetrical waveforms, on the other hand, contain
even-numbered harmonics (which may or may not be accompanied by odd-numbered harmonics as
well).
In the interest of reducing total harmonic content in power systems, the fewer and less diverse the
harmonics, the better one more reason why individual SCRs are favored over TRIACs for complex,
high-power control circuits. One way to make the TRIACs current waveform more symmetrical is
to use a device external to the TRIAC to time the triggering pulse. A DIAC placed in series with
the gate does a fair job of this:

7.6.

THE TRIAC

309

Lamp

AC
source
DIAC breakover voltages tend to be much more symmetrical (the same in one polarity as the
other) than TRIAC triggering voltage thresholds. Since the DIAC prevents any gate current until
the triggering voltage has reached a certain, repeatable level in either direction, the firing point of
the TRIAC from one half-cycle to the next tends to be more consistent, and the waveform more
symmetrical above and below its centerline.
Practically all the characteristics and ratings of SCRs apply equally to TRIACs, except that
TRIACs of course are bidirectional (can handle current in both directions). Not much more needs
to be said about this device except for an important caveat concerning its terminal designations.
From the equivalent circuit diagram shown earlier, one might think that main terminals 1 and 2
were interchangeable. They are not! Although it is helpful to imagine the TRIAC as being composed
of two SCRs joined together, it in fact is constructed from a single piece of semiconducting material,
appropriately doped and layered. The actual operating characteristics may differ slightly from that
of the equivalent model.
This is made most evident by contrasting two simple circuit designs, one that works and one that
doesnt. The following two circuits are a variation of the lamp dimmer circuit shown earlier, the
phase-shifting capacitor and DIAC removed for simplicitys sake. Although the resulting circuit lacks
the fine control ability of the more complex version (with capacitor and DIAC), it does function:

Lamp

AC
source
Suppose we were to swap the two main terminals of the TRIAC around. According to the
equivalent circuit diagram shown earlier in this section, the swap should make no difference. The
circuit ought to work:

This circuit will not work!


Lamp

AC
source

310

CHAPTER 7. THYRISTORS

However, if this circuit is built, it will be found that it does not work! The load will receive no
power, the TRIAC refusing to fire at all, no matter how low or high a resistance value the control
resistor is set to. The key to successfully triggering a TRIAC is to make sure the gate receives its
triggering current from the main terminal 2 side of the circuit (the main terminal on the opposite
side of the TRIAC symbol from the gate terminal). Identification of the MT 1 and MT2 terminals
must be done via the TRIACs part number with reference to a data sheet or book.
REVIEW:
A TRIAC acts much like two SCRs connected back-to-back for bidirectional (AC) operation.
TRIAC controls are more often seen in simple, low-power circuits than complex, high-power
circuits. In large power control circuits, multiple SCRs tend to be favored.
When used to control AC power to a load, TRIACs are often accompanied by DIACs connected
in series with their gate terminals. The DIAC helps the TRIAC fire more symmetrically (more
consistently from one polarity to another).
Main terminals 1 and 2 on a TRIAC are not interchangeable.
To successfully trigger a TRIAC, gate current must come from the main terminal 2 (MT 2 )
side of the circuit!

7.7

Optothyristors

Like bipolar transistors, SCRs and TRIACs are also manufactured as light-sensitive devices, the
action of impinging light replacing the function of triggering voltage.
Optically-controlled SCRs are often known by the acronym LASCR, or Light Activated SCR.
Its symbol, not surprisingly, looks like this:

Light Activated SCR

LASCR
Optically-controlled TRIACs dont receive the honor of having their own acronym, but instead
are humbly known as opto-TRIACs. Their schematic symbol looks like this:

Opto-TRIAC

Optothyristors (a general term for either the LASCR or the opto-TRIAC) are commonly found
inside sealed optoisolator modules.

7.8.

THE UNIJUNCTION TRANSISTOR (UJT) PENDING

7.8

311

The Unijunction Transistor (UJT) PENDING

Programmable Unijunction Transistors (PUTs).

REVIEW:

7.9

The Silicon-Controlled Switch (SCS)

If we take the equivalent circuit for an SCR and add another external terminal, connected to the
base of the top transistor and the collector of the bottom transistor, we have a device known as a
silicon-controlled-switch, or SCS :

The Silicon-Controlled Switch (SCS)

Anode

Anode
P

Cathode
Gate

N
P

N
P
N

Cathode
Physical diagram

Anode
Gate

Anode
Gate
Cathode
Gate

Anode
Anode
Gate
Cathode
Gate
Cathode

Cathode
Equivalent schematic

Schematic symbol

This extra terminal allows more control to be exerted over the device, particularly in the mode
of forced commutation, where an external signal forces it to turn off while the main current through
the device has not yet fallen below the holding current value. Consider the following circuit:

312

CHAPTER 7. THYRISTORS

Equivalent schematic
with two transistors

DC motor start/stop
circuit using an SCS
Motor
R1

Motor
R1

R2

R2

SCS

on

on
SCS
off

off

When the on pushbutton switch is actuated, there is a voltage applied between the cathode
gate and the cathode, forward-biasing the lower transistors base-emitter junction, and turning it
on. The top transistor of the SCS is ready to conduct, having been supplied with a current path
from its emitter terminal (the SCSs anode terminal) through resistor R2 to the positive side of the
power supply. As in the case of the SCR, both transistors turn on and maintain each other in the
on mode. When the lower transistor turns on, it conducts the motors load current, and the motor
starts and runs.
The motor may be stopped by interrupting the power supply, as with an SCR, and this is called
natural commutation. However, the SCS provides us with another means of turning off: forced
commutation by shorting the anode terminal to the cathode. If this is done (by actuating the off
pushbutton switch), the upper transistor within the SCS will lose its emitter current, thus halting
current through the base of the lower transistor. When the lower transistor turns off, it breaks the
circuit for base current through the top transistor (securing its off state), and the motor (making
it stop). The SCS will remain in the off condition until such time that the on pushbutton switch
is re-actuated.
REVIEW:
A silicon-controlled switch, or SCS, is essentially an SCR with an extra gate terminal.
Typically, the load current through an SCS is carried by the anode gate and cathode terminals,
with the cathode gate and anode terminals sufficing as control leads.
An SCS is turned on by applying a positive voltage between the cathode gate and cathode
terminals. It may be turned off (forced commutation) by applying a negative voltage between
the anode and cathode terminals, or simply by shorting those two terminals together. The
anode terminal must be kept positive with respect to the cathode in order for the SCS to
latch.

7.10.

7.10

FIELD-EFFECT-CONTROLLED THYRISTORS

313

Field-effect-controlled thyristors

Two relatively recent technologies designed to reduce the driving (gate trigger current) requirements of classic thyristor devices are the MOS-gated thyristor and the MOS Controlled Thyristor,
or MCT.
The MOS-gated thyristor uses a MOSFET to initiate conduction through the upper (PNP)
transistor of a normal thyristor structure, thus triggering the device. Since a MOSFET requires
negligible current to drive (cause it to saturate), this makes the thyristor as a whole very easy to
trigger:

MOS-gated thyristor
equivalent circuit

Anode

Gate
Cathode

Given the fact that ordinary SCRs are quite easy to drive as it is, the practical advantage of
using an even more sensitive device (a MOSFET) to initiate triggering is debatable. Also, placing
a MOSFET at the gate input of the thyristor now makes it impossible to turn it off by a reversetriggering signal. Only low-current dropout can make this device stop conducting after it has been
latched.
A device of arguably greater value would be a fully-controllable thyristor, whereby a small gate
signal could both trigger the thyristor and force it to turn off. Such a device does exist, and it is
called the MOS Controlled Thyristor, or MCT. It uses a pair of MOSFETs connected to a common
gate terminal, one to trigger the thyristor and the other to untrigger it:

314

CHAPTER 7. THYRISTORS

MOS Controlled Thyristor


(MCT) equivalent circuit

Anode

Gate

Cathode
A positive gate voltage (with respect to the cathode) turns on the upper (N-channel) MOSFET,
allowing base current through the upper (PNP) transistor, which latches the transistor pair in an
on state. Once both transistors are fully latched, there will be little voltage dropped between
anode and cathode, and the thyristor will remain latched so long as the controlled current exceeds
the minimum (holding) current value. However, if a negative gate voltage is applied (with respect
to the anode, which is at nearly the same voltage as the cathode in the latched state), the lower
MOSFET will turn on and short between the lower (NPN) transistors base and emitter terminals,
thus forcing it into cutoff. Once the NPN transistor cuts off, the PNP transistor will drop out of
conduction, and the whole thyristor turns off. Gate voltage has full control over conduction through
the MCT: to turn it on and to turn it off.
This device is still a thyristor, though. If there is zero voltage applied between gate and cathode,
neither MOSFET will turn on. Consequently, the bipolar transistor pair will remain in whatever
state it was last in (hysteresis). So, a brief positive pulse to the gate turns the MCT on, a brief
negative pulse forces it off, and no applied gate voltage lets it remain in whatever state it is already
in. In essence, the MCT is a latching version of the IGBT (Insulated Gate Bipolar Transistor).
REVIEW:
A MOS-gated thyristor uses an N-channel MOSFET to trigger a thyristor, resulting in an
extremely low gate current requirement.
A MOS Controlled Thyristor, or MCT, uses two MOSFETS to exert full control over the
thyristor. A positive gate voltage triggers the device, while a negative gate voltage forces it to
turn off. Zero gate voltage allows the thyristor to remain in whatever state it was previously
in (off, or latched on).

Chapter 8

OPERATIONAL AMPLIFIERS
Contents
8.1
8.2
8.3
8.4
8.5
8.6
8.7
8.8
8.9
8.10
8.11
8.12
8.13

Introduction . . . . . . . . . . . . . . . . . . .
Single-ended and differential amplifiers . . .
The operational amplifier . . . . . . . . . .
Negative feedback . . . . . . . . . . . . . . . .
Divided feedback . . . . . . . . . . . . . . . .
An analogy for divided feedback . . . . . . .
Voltage-to-current signal conversion . . . . .
Averager and summer circuits . . . . . . . .
Building a differential amplifier . . . . . . .
The instrumentation amplifier . . . . . . . .
Differentiator and integrator circuits . . . .
Positive feedback . . . . . . . . . . . . . . . .
Practical considerations . . . . . . . . . . . .
8.13.1 Common-mode gain . . . . . . . . . . . . . .
8.13.2 Offset voltage . . . . . . . . . . . . . . . . .
8.13.3 Bias current . . . . . . . . . . . . . . . . . .
8.13.4 Drift . . . . . . . . . . . . . . . . . . . . . .
8.13.5 Frequency response . . . . . . . . . . . . . .
8.13.6 Input to output phase shift . . . . . . . . . .
8.14 Operational amplifier models . . . . . . . . .
8.15 Data . . . . . . . . . . . . . . . . . . . . . . . .

8.1

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . .
. . . . . . .

315
316
320
326
329
332
337
338
341
343
344
347
351
352
355
357
362
363
364
365
371

Introduction

The operational amplifier is arguably the most useful single device in analog electronic circuitry.
With only a handful of external components, it can be made to perform a wide variety of analog
315

316

CHAPTER 8. OPERATIONAL AMPLIFIERS

signal processing tasks. It is also quite affordable, most general-purpose amplifiers selling for under
a dollar apiece. Modern designs have been engineered with durability in mind as well: several opamps are manufactured that can sustain direct short-circuits on their outputs without damage.
One key to the usefulness of these little circuits is in the engineering principle of feedback,
particularly negative feedback, which constitutes the foundation of almost all automatic control
processes. The principles presented here in operational amplifier circuits, therefore, extend well
beyond the immediate scope of electronics. It is well worth the electronics students time to learn
these principles and learn them well.

8.2

Single-ended and differential amplifiers

For ease of drawing complex circuit diagrams, electronic amplifiers are often symbolized by a simple
triangle shape, where the internal components are not individually represented. This symbology is
very handy for cases where an amplifiers construction is irrelevant to the greater function of the
overall circuit, and it is worthy of familiarization:

General amplifier circuit symbol


+Vsupply
Input

Output
-Vsupply

The +V and -V connections denote the positive and negative sides of the DC power supply,
respectively. The input and output voltage connections are shown as single conductors, because it is
assumed that all signal voltages are referenced to a common connection in the circuit called ground.
Often (but not always!), one pole of the DC power supply, either positive or negative, is that ground
reference point. A practical amplifier circuit (showing the input voltage source, load resistance, and
power supply) might look like this:

+V
Input

+
Output

30 V

Vinput
-V

Rload

Without having to analyze the actual transistor design of the amplifier, you can readily discern
the whole circuits function: to take an input signal (Vin ), amplify it, and drive a load resistance
(Rload ). To complete the above schematic, it would be good to specify the gains of that amplifier
(AV , AI , AP ) and the Q (bias) point for any needed mathematical analysis.

8.2.

SINGLE-ENDED AND DIFFERENTIAL AMPLIFIERS

317

If it is necessary for an amplifier to be able to output true AC voltage (reversing polarity) to


the load, a split DC power supply may be used, whereby the ground point is electrically centered
between +V and -V. Sometimes the split power supply configuration is referred to as a dual power
supply.

+V
Input

+
Output

15 V

Vinput
Rload
-V

15 V
The amplifier is still being supplied with 30 volts overall, but with the split voltage DC power
supply, the output voltage across the load resistor can now swing from a theoretical maximum of
+15 volts to -15 volts, instead of +30 volts to 0 volts. This is an easy way to get true alternating
current (AC) output from an amplifier without resorting to capacitive or inductive (transformer)
coupling on the output. The peak-to-peak amplitude of this amplifiers output between cutoff and
saturation remains unchanged.
By signifying a transistor amplifier within a larger circuit with a triangle symbol, we ease the
task of studying and analyzing more complex amplifiers and circuits. One of these more complex
amplifier types that well be studying is called the differential amplifier. Unlike normal amplifiers,
which amplify a single input signal (often called single-ended amplifiers), differential amplifiers
amplify the voltage difference between two input signals. Using the simplified triangle amplifier
symbol, a differential amplifier looks like this:

Differential amplifier
+Vsupply
Input1

Input2

Output
-Vsupply

The two input leads can be seen on the left-hand side of the triangular amplifier symbol, the
output lead on the right-hand side, and the +V and -V power supply leads on top and bottom. As
with the other example, all voltages are referenced to the circuits ground point. Notice that one
input lead is marked with a (-) and the other is marked with a (+). Because a differential amplifier
amplifies the difference in voltage between the two inputs, each input influences the output voltage
in opposite ways. Consider the following table of input/output voltages for a differential amplifier
with a voltage gain of 4:

318

CHAPTER 8. OPERATIONAL AMPLIFIERS


(-) Input1

2.5

-3

-2

(+) Input2

2.5

-7

Output

10

28

-4

-10

-28

24

-20

Voltage output equation: Vout = AV(Input2 - Input1)


or
Vout = AV(Input(+) - Input(-))

An increasingly positive voltage on the (+) input tends to drive the output voltage more positive,
and an increasingly positive voltage on the (-) input tends to drive the output voltage more negative.
Likewise, an increasingly negative voltage on the (+) input tends to drive the output negative as
well, and an increasingly negative voltage on the (-) input does just the opposite. Because of this
relationship between inputs and polarities, the (-) input is commonly referred to as the inverting
input and the (+) as the noninverting input.
It may be helpful to think of a differential amplifier as a variable voltage source controlled by a
sensitive voltmeter, as such:

+V

G
+
+

-V

Bear in mind that the above illustration is only a model to aid in understanding the behavior of a
differential amplifier. It is not a realistic schematic of its actual design. The G symbol represents
a galvanometer, a sensitive voltmeter movement. The potentiometer connected between +V and -V
provides a variable voltage at the output pin (with reference to one side of the DC power supply),
that variable voltage set by the reading of the galvanometer. It must be understood that any load
powered by the output of a differential amplifier gets its current from the DC power source (battery),
not the input signal. The input signal (to the galvanometer) merely controls the output.
This concept may at first be confusing to students new to amplifiers. With all these polarities
and polarity markings (- and +) around, its easy to get confused and not know what the output of
a differential amplifier will be. To address this potential confusion, heres a simple rule to remember:

8.2.

SINGLE-ENDED AND DIFFERENTIAL AMPLIFIERS

Differential
input voltage

Differential
input voltage

Output
- voltage

319

Output
voltage

When the polarity of the differential voltage matches the markings for inverting and noninverting
inputs, the output will be positive. When the polarity of the differential voltage clashes with the
input markings, the output will be negative. This bears some similarity to the mathematical sign
displayed by digital voltmeters based on input voltage polarity. The red test lead of the voltmeter
(often called the positive lead because of the color reds popular association with the positive side
of a power supply in electronic wiring) is more positive than the black, the meter will display a
positive voltage figure, and vice versa:

Differential
input voltage

Differential
input voltage

blk

6V
+

red

blk

6V
-

red

+ 6.00 V
Digital Voltmeter

- 6.00 V
Digital Voltmeter

Just as a voltmeter will only display the voltage between its two test leads, an ideal differential
amplifier only amplifies the potential difference between its two input connections, not the voltage
between any one of those connections and ground. The output polarity of a differential amplifier, just
like the signed indication of a digital voltmeter, depends on the relative polarities of the differential
voltage between the two input connections.
If the input voltages to this amplifier represented mathematical quantities (as is the case within
analog computer circuitry), or physical process measurements (as is the case within analog electronic
instrumentation circuitry), you can see how a device such as a differential amplifier could be very
useful. We could use it to compare two quantities to see which is greater (by the polarity of the
output voltage), or perhaps we could compare the difference between two quantities (such as the
level of liquid in two tanks) and flag an alarm (based on the absolute value of the amplifier output) if
the difference became too great. In basic automatic control circuitry, the quantity being controlled
(called the process variable) is compared with a target value (called the setpoint), and decisions
are made as to how to act based on the discrepancy between these two values. The first step in
electronically controlling such a scheme is to amplify the difference between the process variable and

320

CHAPTER 8. OPERATIONAL AMPLIFIERS

the setpoint with a differential amplifier. In simple controller designs, the output of this differential
amplifier can be directly utilized to drive the final control element (such as a valve) and keep the
process reasonably close to setpoint.
REVIEW:
A shorthand symbol for an electronic amplifier is a triangle, the wide end signifying the
input side and the narrow end signifying the output. Power supply lines are often omitted in
the drawing for simplicity.
To facilitate true AC output from an amplifier, we can use what is called a split or dual power
supply, with two DC voltage sources connected in series with the middle point grounded, giving
a positive voltage to ground (+V) and a negative voltage to ground (-V). Split power supplies
like this are frequently used in differential amplifier circuits.
Most amplifiers have one input and one output. Differential amplifiers have two inputs and
one output, the output signal being proportional to the difference in signals between the two
inputs.
The voltage output of a differential amplifier is determined by the following equation: V out =
AV (Vnoninv - Vinv )

8.3

The operational amplifier

Long before the advent of digital electronic technology, computers were built to electronically perform calculations by employing voltages and currents to represent numerical quantities. This was
especially useful for the simulation of physical processes. A variable voltage, for instance, might represent velocity or force in a physical system. Through the use of resistive voltage dividers and voltage
amplifiers, the mathematical operations of division and multiplication could be easily performed on
these signals.
The reactive properties of capacitors and inductors lend themselves well to the simulation of
variables related by calculus functions. Remember how the current through a capacitor was a
function of the voltages rate of change, and how that rate of change was designated in calculus as
the derivative? Well, if voltage across a capacitor were made to represent the velocity of an object,
the current through the capacitor would represent the force required to accelerate or decelerate that
object, the capacitors capacitance representing the objects mass:
iC = C dv
F = m dv
dt
dt
Where,
iC = Instantaneous current
through capacitor
C = Capacitance in farads
dv = Rate of change of
dt
voltage over time

Where,
F = Force applied to object
m = Mass of object
dv = Rate of change of
dt
velocity over time

This analog electronic computation of the calculus derivative function is technically known as
differentiation, and it is a natural function of a capacitors current in relation to the voltage applied

8.3.

THE OPERATIONAL AMPLIFIER

321

across it. Note that this circuit requires no programming to perform this relatively advanced
mathematical function as a digital computer would.

Electronic circuits are very easy and inexpensive to create compared to complex physical systems,
so this kind of analog electronic simulation was widely used in the research and development of
mechanical systems. For realistic simulation, though, amplifier circuits of high accuracy and easy
configurability were needed in these early computers.

It was found in the course of analog computer design that differential amplifiers with extremely
high voltage gains met these requirements of accuracy and configurability better than single-ended
amplifiers with custom-designed gains. Using simple components connected to the inputs and output
of the high-gain differential amplifier, virtually any gain and any function could be obtained from the
circuit, overall, without adjusting or modifying the internal circuitry of the amplifier itself. These
high-gain differential amplifiers came to be known as operational amplifiers, or op-amps, because of
their application in analog computers mathematical operations.

Modern op-amps, like the popular model 741, are high-performance, inexpensive integrated circuits. Their input impedances are quite high, the inputs drawing currents in the range of half
a microamp (maximum) for the 741, and far less for op-amps utilizing field-effect input transistors. Output impedance is typically quite low, about 75 for the model 741, and many models
have built-in output short circuit protection, meaning that their outputs can be directly shorted
to ground without causing harm to the internal circuitry. With direct coupling between op-amps
internal transistor stages, they can amplify DC signals just as well as AC (up to certain maximum
voltage-risetime limits). It would cost far more in money and time to design a comparable discretetransistor amplifier circuit to match that kind of performance, unless high power capability was
required. For these reasons, op-amps have all but obsoleted discrete-transistor signal amplifiers in
many applications.

The following diagram shows the pin connections for single op-amps (741 included) when housed
in an 8-pin DIP (Dual Inline Package) integrated circuit:

322

CHAPTER 8. OPERATIONAL AMPLIFIERS

Typical 8-pin "DIP" op-amp


integrated circuit
No
+V
connection

Offset
Output null

Offset
null

-V

Some models of op-amp come two to a package, including the popular models TL082 and 1458.
These are called dual units, and are typically housed in an 8-pin DIP package as well, with the
following pin connections:

Dual op-amp in 8-pin DIP


+V
8

+
+

4
-V

Operational amplifiers are also available four to a package, usually in 14-pin DIP arrangements.
Unfortunately, pin assignments arent as standard for these quad op-amps as they are for the
dual or single units. Consult the manufacturer datasheet(s) for details.
Practical operational amplifier voltage gains are in the range of 200,000 or more, which makes

8.3.

THE OPERATIONAL AMPLIFIER

323

them almost useless as an analog differential amplifier by themselves. For an op-amp with a voltage
gain (AV ) of 200,000 and a maximum output voltage swing of +15V/-15V, all it would take is a
differential input voltage of 75 V (microvolts) to drive it to saturation or cutoff! Before we take
a look at how external components are used to bring the gain down to a reasonable level, lets
investigate applications for the bare op-amp by itself.

One application is called the comparator. For all practical purposes, we can say that the output
of an op-amp will be saturated fully positive if the (+) input is more positive than the (-) input,
and saturated fully negative if the (+) input is less positive than the (-) input. In other words,
an op-amps extremely high voltage gain makes it useful as a device to compare two voltages and
change output voltage states when one input exceeds the other in magnitude.

+V
Vin

LED

-V

In the above circuit, we have an op-amp connected as a comparator, comparing the input voltage
with a reference voltage set by the potentiometer (R1 ). If Vin drops below the voltage set by R1 ,
the op-amps output will saturate to +V, thereby lighting up the LED. Otherwise, if V in is above
the reference voltage, the LED will remain off. If Vin is a voltage signal produced by a measuring
instrument, this comparator circuit could function as a low alarm, with the trip-point set by R 1 .
Instead of an LED, the op-amp output could drive a relay, a transistor, an SCR, or any other device
capable of switching power to a load such as a solenoid valve, to take action in the event of a low
alarm.

Another application for the comparator circuit shown is a square-wave converter. Suppose that
the input voltage applied to the inverting (-) input was an AC sine wave rather than a stable DC
voltage. In that case, the output voltage would transition between opposing states of saturation
whenever the input voltage was equal to the reference voltage produced by the potentiometer. The
result would be a square wave:

324

CHAPTER 8. OPERATIONAL AMPLIFIERS

+V

Vin

Vout

-V
Vin

Vout

Adjustments to the potentiometer setting would change the reference voltage applied to the
noninverting (+) input, which would change the points at which the sine wave would cross, changing
the on/off times, or duty cycle of the square wave:

+V

Vin

Vout

-V
Vin

Vout

It should be evident that the AC input voltage would not have to be a sine wave in particular
for this circuit to perform the same function. The input voltage could be a triangle wave, sawtooth
wave, or any other sort of wave that ramped smoothly from positive to negative to positive again.
This sort of comparator circuit is very useful for creating square waves of varying duty cycle. This
technique is sometimes referred to as pulse-width modulation, or PWM (varying, or modulating a
waveform according to a controlling signal, in this case the signal produced by the potentiometer).
Another comparator application is that of the bargraph driver. If we had several op-amps
connected as comparators, each with its own reference voltage connected to the inverting input, but

8.3.

THE OPERATIONAL AMPLIFIER

325

each one monitoring the same voltage signal on their noninverting inputs, we could build a bargraphstyle meter such as what is commonly seen on the face of stereo tuners and graphic equalizers. As the
signal voltage (representing radio signal strength or audio sound level) increased, each comparator
would turn on in sequence and send power to its respective LED. With each comparator switching
on at a different level of audio sound, the number of LEDs illuminated would indicate how strong
the signal was.

+V

Simple bargraph driver circuit

LED4

LED3

LED2

LED1

+
-V

-V
Vin

In the circuit shown above, LED1 would be the first to light up as the input voltage increased in
a positive direction. As the input voltage continued to increase, the other LEDs would illuminate
in succession, until all were lit.
This very same technology is used in some analog-to-digital signal converters, namely the flash
converter, to translate an analog signal quantity into a series of on/off voltages representing a digital
number.
REVIEW:
A triangle shape is the generic symbol for an amplifier circuit, the wide end signifying the
input and the narrow end signifying the output.
Unless otherwise specified, all voltages in amplifier circuits are referenced to a common ground
point, usually connected to one terminal of the power supply. This way, we can speak of
a certain amount of voltage being on a single wire, while realizing that voltage is always
measured between two points.

326

CHAPTER 8. OPERATIONAL AMPLIFIERS


A differential amplifier is one amplifying the voltage difference between two signal inputs. In
such a circuit, one input tends to drive the output voltage to the same polarity of the input
signal, while the other input does just the opposite. Consequently, the first input is called the
noninverting (+) input and the second is called the inverting (-) input.
An operational amplifier (or op-amp for short) is a differential amplifier with an extremely high
voltage gain (AV = 200,000 or more). Its name hails from its original use in analog computer
circuitry (performing mathematical operations).
Op-amps typically have very high input impedances and fairly low output impedances.
Sometimes op-amps are used as signal comparators, operating in full cutoff or saturation mode
depending on which input (inverting or noninverting) has the greatest voltage. Comparators
are useful in detecting greater-than signal conditions (comparing one to the other).
One comparator application is called the pulse-width modulator, and is made by comparing a
sine-wave AC signal against a DC reference voltage. As the DC reference voltage is adjusted,
the square-wave output of the comparator changes its duty cycle (positive versus negative
times). Thus, the DC reference voltage controls, or modulates the pulse width of the output
voltage.

8.4

Negative feedback

If we connect the output of an op-amp to its inverting input and apply a voltage signal to the
noninverting input, we find that the output voltage of the op-amp closely follows that input voltage
(Ive neglected to draw in the power supply, +V/-V wires, and ground symbol for simplicity):

Vin

Vout

As Vin increases, Vout will increase in accordance with the differential gain. However, as Vout
increases, that output voltage is fed back to the inverting input, thereby acting to decrease the
voltage differential between inputs, which acts to bring the output down. What will happen for any
given voltage input is that the op-amp will output a voltage very nearly equal to V in , but just low
enough so that theres enough voltage difference left between Vin and the (-) input to be amplified
to generate the output voltage.
The circuit will quickly reach a point of stability (known as equilibrium in physics), where the
output voltage is just the right amount to maintain the right amount of differential, which in turn
produces the right amount of output voltage. Taking the op-amps output voltage and coupling
it to the inverting input is a technique known as negative feedback, and it is the key to having a
self-stabilizing system (this is true not only of op-amps, but of any dynamic system in general). This
stability gives the op-amp the capacity to work in its linear (active) mode, as opposed to merely
being saturated fully on or off as it was when used as a comparator, with no feedback at all.
Because the op-amps gain is so high, the voltage on the inverting input can be maintained almost
equal to Vin . Lets say that our op-amp has a differential voltage gain of 200,000. If V in equals

8.4.

NEGATIVE FEEDBACK

327

6 volts, the output voltage will be 5.999970000149999 volts. This creates just enough differential
voltage (6 volts - 5.999970000149999 volts = 29.99985 V) to cause 5.999970000149999 volts to be
manifested at the output terminal, and the system holds there in balance. As you can see, 29.99985
V is not a lot of differential, so for practical calculations, we can assume that the differential voltage
between the two input wires is held by negative feedback exactly at 0 volts.

The effects of negative feedback

29.99985 V

+
5.999970000149999 V

6V

The effects of negative feedback


(rounded figures)

0V

6V

6V

One great advantage to using an op-amp with negative feedback is that the actual voltage gain of
the op-amp doesnt matter, so long as its very large. If the op-amps differential gain were 250,000
instead of 200,000, all it would mean is that the output voltage would hold just a little closer to V in
(less differential voltage needed between inputs to generate the required output). In the circuit just
illustrated, the output voltage would still be (for all practical purposes) equal to the non-inverting
input voltage. Op-amp gains, therefore, do not have to be precisely set by the factory in order for the
circuit designer to build an amplifier circuit with precise gain. Negative feedback makes the system
self-correcting. The above circuit as a whole will simply follow the input voltage with a stable gain
of 1.
Going back to our differential amplifier model, we can think of the operational amplifier as
being a variable voltage source controlled by an extremely sensitive null detector, the kind of meter
movement or other sensitive measurement device used in bridge circuits to detect a condition of

328

CHAPTER 8. OPERATIONAL AMPLIFIERS

balance (zero volts). The potentiometer inside the op-amp creating the variable voltage will move
to whatever position it must to balance the inverting and noninverting input voltages so that the
null detector has zero voltage across it:

+V

6V
0V

null
+

-V

6V

As the potentiometer will move to provide an output voltage necessary to satisfy the null
detector at an indication of zero volts, the output voltage becomes equal to the input voltage:
in this case, 6 volts. If the input voltage changes at all, the potentiometer inside the op-amp will
change position to hold the null detector in balance (indicating zero volts), resulting in an output
voltage approximately equal to the input voltage at all times.
This will hold true within the range of voltages that the op-amp can output. With a power
supply of +15V/-15V, and an ideal amplifier that can swing its output voltage just as far, it will
faithfully follow the input voltage between the limits of +15 volts and -15 volts. For this reason,
the above circuit is known as a voltage follower. Like its one-transistor counterpart, the commoncollector (emitter-follower) amplifier, it has a voltage gain of 1, a high input impedance, a low
output impedance, and a high current gain. Voltage followers are also known as voltage buffers,
and are used to boost the current-sourcing ability of voltage signals too weak (too high of source
impedance) to directly drive a load. The op-amp model shown in the last illustration depicts how
the output voltage is essentially isolated from the input voltage, so that current on the output pin
is not supplied by the input voltage source at all, but rather from the power supply powering the
op-amp.
It should be mentioned that many op-amps cannot swing their output voltages exactly to +V/-V
power supply rail voltages. The model 741 is one of those that cannot: when saturated, its output
voltage peaks within about one volt of the +V power supply voltage and within about 2 volts of
the -V power supply voltage. Therefore, with a split power supply of +15/-15 volts, a 741 op-amps
output may go as high as +14 volts or as low as -13 volts (approximately), but no further. This is
due to its bipolar transistor design. These two voltage limits are known as the positive saturation
voltage and negative saturation voltage, respectively. Other op-amps, such as the model 3130 with
field-effect transistors in the final output stage, have the ability to swing their output voltages within
millivolts of either power supply rail voltage. Consequently, their positive and negative saturation
voltages are practically equal to the supply voltages.
REVIEW:

8.5.

DIVIDED FEEDBACK

329

Connecting the output of an op-amp to its inverting (-) input is called negative feedback. This
term can be broadly applied to any dynamic system where the output signal is fed back to
the input somehow so as to reach a point of equilibrium (balance).
When the output of an op-amp is directly connected to its inverting (-) input, a voltage follower
will be created. Whatever signal voltage is impressed upon the noninverting (+) input will be
seen on the output.
An op-amp with negative feedback will try to drive its output voltage to whatever level necessary so that the differential voltage between the two inputs is practically zero. The higher
the op-amp differential gain, the closer that differential voltage will be to zero.
Some op-amps cannot produce an output voltage equal to their supply voltage when saturated.
The model 741 is one of these. The upper and lower limits of an op-amps output voltage swing
are known as positive saturation voltage and negative saturation voltage, respectively.

8.5

Divided feedback

If we add a voltage divider to the negative feedback wiring so that only a fraction of the output
voltage is fed back to the inverting input instead of the full amount, the output voltage will be a
multiple of the input voltage (please bear in mind that the power supply connections to the op-amp
have been omitted once again for simplicitys sake):

The effects of divided negative feedback


6 mA R1
1 k
6V

0V

R2 6 mA

6V

6V

1 k
12 V

All voltage figures shown in


reference to ground

If R1 and R2 are both equal and Vin is 6 volts, the op-amp will output whatever voltage is needed
to drop 6 volts across R1 (to make the inverting input voltage equal to 6 volts, as well, keeping the
voltage difference between the two inputs equal to zero). With the 2:1 voltage divider of R 1 and R2 ,
this will take 12 volts at the output of the op-amp to accomplish.
Another way of analyzing this circuit is to start by calculating the magnitude and direction of
current through R1 , knowing the voltage on either side (and therefore, by subtraction, the voltage
across R1 ), and R1 s resistance. Since the left-hand side of R1 is connected to ground (0 volts) and
the right-hand side is at a potential of 6 volts (due to the negative feedback holding that point equal
to Vin ), we can see that we have 6 volts across R1 . This gives us 6 mA of current through R1 from

330

CHAPTER 8. OPERATIONAL AMPLIFIERS

left to right. Because we know that both inputs of the op-amp have extremely high impedance,
we can safely assume they wont add or subtract any current through the divider. In other words,
we can treat R1 and R2 as being in series with each other: all of the electrons flowing through R1
must flow through R2 . Knowing the current through R2 and the resistance of R2 , we can calculate
the voltage across R2 (6 volts), and its polarity. Counting up voltages from ground (0 volts) to the
right-hand side of R2 , we arrive at 12 volts on the output.
Upon examining the last illustration, one might wonder, where does that 1 mA of current go?
The last illustration doesnt show the entire current path, but in reality it comes from the negative
side of the DC power supply, through ground, through R1 , through R2 , through the output pin of the
op-amp, and then back to the positive side of the DC power supply through the output transistor(s)
of the op-amp. Using the null detector/potentiometer model of the op-amp, the current path looks
like this:

R1

R2

1 k

1 k
+V

null
+

-V

6V

The 6 volt signal source does not have to supply any current for the circuit: it merely commands
the op-amp to balance voltage between the inverting (-) and noninverting (+) input pins, and in so
doing produce an output voltage that is twice the input due to the dividing effect of the two 1 k
resistors.
We can change the voltage gain of this circuit, overall, just by adjusting the values of R 1 and R2
(changing the ratio of output voltage that is fed back to the inverting input). Gain can be calculated
by the following formula:

AV =

R2
+1
R1

Note that the voltage gain for this design of amplifier circuit can never be less than 1. If we were
to lower R2 to a value of zero ohms, our circuit would be essentially identical to the voltage follower,
with the output directly connected to the inverting input. Since the voltage follower has a gain of
1, this sets the lower gain limit of the noninverting amplifier. However, the gain can be increased
far beyond 1, by increasing R2 in proportion to R1 .
Also note that the polarity of the output matches that of the input, just as with a voltage
follower. A positive input voltage results in a positive output voltage, and vice versa (with respect

8.5.

DIVIDED FEEDBACK

331

to ground). For this reason, this circuit is referred to as a noninverting amplifier.


Just as with the voltage follower, we see that the differential gain of the op-amp is irrelevant, so
long as its very high. The voltages and currents in this circuit would hardly change at all if the
op-amps voltage gain were 250,000 instead of 200,000. This stands as a stark contrast to singletransistor amplifier circuit designs, where the Beta of the individual transistor greatly influenced
the overall gains of the amplifier. With negative feedback, we have a self-correcting system that
amplifies voltage according to the ratios set by the feedback resistors, not the gains internal to the
op-amp.
Lets see what happens if we retain negative feedback through a voltage divider, but apply the
input voltage at a different location:

6 mA R1
1 k
6V

R2 6 mA

0V

0V

1 k
-6 V

All voltage figures shown in


reference to ground

By grounding the noninverting input, the negative feedback from the output seeks to hold the
inverting inputs voltage at 0 volts, as well. For this reason, the inverting input is referred to in
this circuit as a virtual ground, being held at ground potential (0 volts) by the feedback, yet not
directly connected to (electrically common with) ground. The input voltage this time is applied to
the left-hand end of the voltage divider (R1 = R2 = 1 k again), so the output voltage must swing
to -6 volts in order to balance the middle at ground potential (0 volts). Using the same techniques
as with the noninverting amplifier, we can analyze this circuits operation by determining current
magnitudes and directions, starting with R1 , and continuing on to determining the output voltage.
We can change the overall voltage gain of this circuit, overall, just by adjusting the values of R 1
and R2 (changing the ratio of output voltage that is fed back to the inverting input). Gain can be
calculated by the following formula:

AV =

R2
R1

Note that this circuits voltage gain can be less than 1, depending solely on the ratio of R 2 to R1 .
Also note that the output voltage is always the opposite polarity of the input voltage. A positive
input voltage results in a negative output voltage, and vice versa (with respect to ground). For this
reason, this circuit is referred to as an inverting amplifier. Sometimes, the gain formula contains a
negative sign (before the R2 /R1 fraction) to reflect this reversal of polarities.
These two amplifier circuits weve just investigated serve the purpose of multiplying or dividing
the magnitude of the input voltage signal. This is exactly how the mathematical operations of
multiplication and division are typically handled in analog computer circuitry.
REVIEW:
By connecting the inverting (-) input of an op-amp directly to the output, we get negative
feedback, which gives us a voltage follower circuit. By connecting that negative feedback

332

CHAPTER 8. OPERATIONAL AMPLIFIERS


through a resistive voltage divider (feeding back a fraction of the output voltage to the inverting
input), the output voltage becomes a multiple of the input voltage.
A negative-feedback op-amp circuit with the input signal going to the noninverting (+) input
is called a noninverting amplifier. The output voltage will be the same polarity as the input.
Voltage gain is given by the following equation: AV = (R2 /R1 ) + 1
A negative-feedback op-amp circuit with the input signal going to the bottom of the resistive
voltage divider, with the noninverting (+) input grounded, is called an inverting amplifier. Its
output voltage will be the opposite polarity of the input. Voltage gain is given by the following
equation: AV = R2 /R1

8.6

An analogy for divided feedback

A helpful analogy for understanding divided feedback amplifier circuits is that of a mechanical lever,
with relative motion of the levers ends representing change in input and output voltages, and the
fulcrum (pivot point) representing the location of the ground point, real or virtual.
Take for example the following noninverting op-amp circuit. We know from the prior section
that the voltage gain of a noninverting amplifier configuration can never be less than unity (1). If
we draw a lever diagram next to the amplifier schematic, with the distance between fulcrum and
lever ends representative of resistor values, the motion of the lever will signify changes in voltage at
the input and output terminals of the amplifier:

Vout
R2

R1

Vout = 2(Vin)

Vin
R1

R2

1 k

0V

1 k

Vout

Vin

Physicists call this type of lever, with the input force (effort) applied between the fulcrum and
output (load), a third-class lever. It is characterized by an output displacement (motion) at least
as large than the input displacement a gain of at least 1 and in the same direction. Applying

8.6.

AN ANALOGY FOR DIVIDED FEEDBACK

333

a positive input voltage to this op-amp circuit is analogous to displacing the input point on the
lever upward:

Vout

Vin
Vout = 2(Vin)
R1

R2

1 k

1 k

0V

Vout
-

+
Vin
-

Due to the displacement-amplifying characteristics of the lever, the output point will move
twice as far as the input point, and in the same direction. In the electronic circuit, the output
voltage will equal twice the input, with the same polarity. Applying a negative input voltage is
analogous to moving the lever downward from its level zero position, resulting in an amplified
output displacement that is also negative:

334

CHAPTER 8. OPERATIONAL AMPLIFIERS

Vin
Vout
R1

R2

1 k

1 k

0V

+
Vin

Vout = 2(Vin)

Vout
+

If we alter the resistor ratio R2 /R1 , we change the gain of the op-amp circuit. In lever terms,
this means moving the input point in relation to the fulcrum and lever end, which similarly changes
the displacement gain of the machine:

Vout

R1

R2
Vout = 4(Vin)

Vin
R1

R2

1 k

0V

3 k

Vout

Vin

Now, any input signal will become amplified by a factor of four instead of by a factor of two:

8.6.

AN ANALOGY FOR DIVIDED FEEDBACK

335

Vout
Vin
Vout = 4(Vin)
R1

R2

1 k

0V

3 k
+
Vout
-

+
+
Vin
-

Inverting op-amp circuits may be modeled using the lever analogy as well. With the inverting
configuration, the ground point of the feedback voltage divider is the op-amps inverting input with
the input to the left and the output to the right. This is mechanically equivalent to a first-class
lever, where the input force (effort) is on the opposite side of the fulcrum from the output (load):

Vin
Vin

R1

R2

R1

R2

1 k

Vout
Vout = -(Vin)

1 k
Vout

With equal-value resistors (equal-lengths of lever on each side of the fulcrum), the output voltage
(displacement) will be equal in magnitude to the input voltage (displacement), but of the opposite
polarity (direction). A positive input results in a negative output:

336

CHAPTER 8. OPERATIONAL AMPLIFIERS

Vout

Vin
R1
+
Vin

1 k

R2

Vout = -(Vin)

1 k
Vout
+

Changing the resistor ratio R2 /R1 changes the gain of the amplifier circuit, just as changing the
fulcrum position on the lever changes its mechanical displacement gain. Consider the following
example, where R2 is made twice as large as R1 :

Vout

Vin

+
Vin
-

R1

R2

1 k

2 k

Vout = -2(Vin)
Vout
+

With the inverting amplifier configuration, though, gains of less than 1 are possible, just as with
first-class levers. Reversing R2 and R1 values is analogous to moving the fulcrum to its complementary position on the lever: one-third of the way from the output end. There, the output displacement
will be one-half the input displacement:

8.7.

VOLTAGE-TO-CURRENT SIGNAL CONVERSION

337

Vout
Vin

+
Vin
-

R1

R2

2 k

1 k

8.7

Vout = -0.5(Vin)
Vout
+

Voltage-to-current signal conversion

In instrumentation circuitry, DC signals are often used as analog representations of physical measurements such as temperature, pressure, flow, weight, and motion. Most commonly, DC current
signals are used in preference to DC voltage signals, because current signals are exactly equal in
magnitude throughout the series circuit loop carrying current from the source (measuring device)
to the load (indicator, recorder, or controller), whereas voltage signals in a parallel circuit may vary
from one end to the other due to resistive wire losses. Furthermore, current-sensing instruments typically have low impedances (while voltage-sensing instruments have high impedances), which gives
current-sensing instruments greater electrical noise immunity.
In order to use current as an analog representation of a physical quantity, we have to have some
way of generating a precise amount of current within the signal circuit. But how do we generate
a precise current signal when we might not know the resistance of the loop? The answer is to use
an amplifier designed to hold current to a prescribed value, applying as much or as little voltage as
necessary to the load circuit to maintain that value. Such an amplifier performs the function of a
current source. An op-amp with negative feedback is a perfect candidate for such a task:

338

CHAPTER 8. OPERATIONAL AMPLIFIERS

4 to 20 mA

250

Rload
+

4 to 20 mA

+
Vin

1 to 5 volt signal range


-

The input voltage to this circuit is assumed to be coming from some type of physical transducer/amplifier arrangement, calibrated to produce 1 volt at 0 percent of physical measurement,
and 5 volts at 100 percent of physical measurement. The standard analog current signal range is 4
mA to 20 mA, signifying 0% to 100% of measurement range, respectively. At 5 volts input, the 250
(precision) resistor will have 5 volts applied across it, resulting in 20 mA of current in the large loop
circuit (with Rload ). It does not matter what resistance value Rload is, or how much wire resistance
is present in that large loop, so long as the op-amp has a high enough power supply voltage to
output the voltage necessary to get 20 mA flowing through Rload . The 250 resistor establishes the
relationship between input voltage and output current, in this case creating the equivalence of 1-5
V in / 4-20 mA out. If we were converting the 1-5 volt input signal to a 10-50 mA output signal (an
older, obsolete instrumentation standard for industry), wed use a 100 precision resistor instead.
Another name for this circuit is transconductance amplifier. In electronics, transconductance is
the mathematical ratio of current change divided by voltage change (I / V), and it is measured
in the unit of Siemens, the same unit used to express conductance (the mathematical reciprocal of
resistance: current/voltage). In this circuit, the transconductance ratio is fixed by the value of the
250 resistor, giving a linear current-out/voltage-in relationship.
REVIEW:
In industry, DC current signals are often used in preference to DC voltage signals as analog
representations of physical quantities. Current in a series circuit is absolutely equal at all
points in that circuit regardless of wiring resistance, whereas voltage in a parallel-connected
circuit may vary from end to end because of wire resistance, making current-signaling more
accurate from the transmitting to the receiving instrument.
Voltage signals are relatively easy to produce directly from transducer devices, whereas accurate current signals are not. Op-amps can be used to convert a voltage signal into a current
signal quite easily. In this mode, the op-amp will output whatever voltage is necessary to
maintain current through the signaling circuit at the proper value.

8.8

Averager and summer circuits

If we take three equal resistors and connect one end of each to a common point, then apply three
input voltages (one to each of the resistors free ends), the voltage seen at the common point will be

8.8.

AVERAGER AND SUMMER CIRCUITS

339

the mathematical average of the three.

"Passive averager" circuit


R1
R2
Vout =
R3
V1

V2

V3

V1
V2
V3
R1 + R2 + R3
1
1
1
R1 + R2 + R3

With equal value resistors:


V1 + V2 + V3
Vout =
3

This circuit is really nothing more than a practical application of Millmans Theorem:

R1

R2

R3
Vout =

V1

V2

V3

V1
V2
V3
R1 + R2 + R3
1
1
1
R1 + R2 + R3

This circuit is commonly known as a passive averager, because it generates an average voltage
with non-amplifying components. Passive simply means that it is an unamplified circuit. The
large equation to the right of the averager circuit comes from Millmans Theorem, which describes
the voltage produced by multiple voltage sources connected together through individual resistances.
Since the three resistors in the averager circuit are equal to each other, we can simplify Millmans
formula by writing R1 , R2 , and R3 simply as R (one, equal resistance instead of three individual
resistances):

340

CHAPTER 8. OPERATIONAL AMPLIFIERS

Vout =

V1
V2
V3
+
+
R
R
R
1
1
1
R + R + R

Vout =

V1 + V2 + V3
R
3
R

Vout =

V1 + V2 + V3
3

If we take a passive averager and use it to connect three input voltages into an op-amp amplifier
circuit with a gain of 3, we can turn this averaging function into an addition function. The result is
called a noninverting summer circuit:

1 k

V1
V2
V3

2 k

Vout

With a voltage divider composed of a 2 k / 1 k combination, the noninverting amplifier circuit


will have a voltage gain of 3. By taking the voltage from the passive averager, which is the sum
of V1 , V2 , and V3 divided by 3, and multiplying that average by 3, we arrive at an output voltage
equal to the sum of V1 , V2 , and V3 :

Vout = 3

V1 + V2 + V3
3

Vout = V1 + V2 + V3
Much the same can be done with an inverting op-amp amplifier, using a passive averager as part
of the voltage divider feedback circuit. The result is called an inverting summer circuit:

8.9.

BUILDING A DIFFERENTIAL AMPLIFIER

V1
V2
V3

I1

R
R

341

0V
I2
I3

I1 + I2 + I3

Vout

0V
Now, with the right-hand sides of the three averaging resistors connected to the virtual ground
point of the op-amps inverting input, Millmans Theorem no longer directly applies as it did before.
The voltage at the virtual ground is now held at 0 volts by the op-amps negative feedback, whereas
before it was free to float to the average value of V1 , V2 , and V3 . However, with all resistor
values equal to each other, the currents through each of the three resistors will be proportional to
their respective input voltages. Since those three currents will add at the virtual ground node, the
algebraic sum of those currents through the feedback resistor will produce a voltage at V out equal
to V1 + V2 + V3 , except with reversed polarity. The reversal in polarity is what makes this circuit
an inverting summer:

Vout = -(V1 + V2 + V3)


Summer (adder) circuits are quite useful in analog computer design, just as multiplier and divider
circuits would be. Again, it is the extremely high differential gain of the op-amp which allows us to
build these useful circuits with a bare minimum of components.

REVIEW:

A summer circuit is one that sums, or adds, multiple analog voltage signals together. There
are two basic varieties of op-amp summer circuits: noninverting and inverting.

8.9

Building a differential amplifier

An op-amp with no feedback is already a differential amplifier, amplifying the voltage difference
between the two inputs. However, its gain cannot be controlled, and it is generally too high to be
of any practical use. So far, our application of negative feedback to op-amps has resulting in the
practical loss of one of the inputs, the resulting amplifier only good for amplifying a single voltage
signal input. With a little ingenuity, however, we can construct an op-amp circuit maintaining both
voltage inputs, yet with a controlled gain set by external resistors.

342

CHAPTER 8. OPERATIONAL AMPLIFIERS

V1

Vout

+
R

V2
If all the resistor values are equal, this amplifier will have a differential voltage gain of 1. The
analysis of this circuit is essentially the same as that of an inverting amplifier, except that the
noninverting input (+) of the op-amp is at a voltage equal to a fraction of V 2 , rather than being
connected directly to ground. As would stand to reason, V2 functions as the noninverting input and
V1 functions as the inverting input of the final amplifier circuit. Therefore:

Vout = V2 - V1
If we wanted to provide a differential gain of anything other than 1, we would have to adjust
the resistances in both upper and lower voltage dividers, necessitating multiple resistor changes and
balancing between the two dividers for symmetrical operation. This is not always practical, for
obvious reasons.
Another limitation of this amplifier design is the fact that its input impedances are rather low
compared to that of some other op-amp configurations, most notably the noninverting (singleended input) amplifier. Each input voltage source has to drive current through a resistance, which
constitutes far less impedance than the bare input of an op-amp alone. The solution to this problem,
fortunately, is quite simple. All we need to do is buffer each input voltage signal through a voltage
follower like this:

V1

Vout

V2

Now the V1 and V2 input lines are connected straight to the inputs of two voltage-follower opamps, giving very high impedance. The two op-amps on the left now handle the driving of current
through the resistors instead of letting the input voltage sources (whatever they may be) do it. The
increased complexity to our circuit is minimal for a substantial benefit.

8.10.

8.10

THE INSTRUMENTATION AMPLIFIER

343

The instrumentation amplifier

As suggested before, it is beneficial to be able to adjust the gain of the amplifier circuit without
having to change more than one resistor value, as is necessary with the previous design of differential
amplifier. The so-called instrumentation builds on the last version of differential amplifier to give
us that capability:

V1

R
1

Rgain

V2

Vout

This intimidating circuit is constructed from a buffered differential amplifier stage with three
new resistors linking the two buffer circuits together. Consider all resistors to be of equal value
except for Rgain . The negative feedback of the upper-left op-amp causes the voltage at point 1 (top
of Rgain ) to be equal to V1 . Likewise, the voltage at point 2 (bottom of Rgain ) is held to a value
equal to V2 . This establishes a voltage drop across Rgain equal to the voltage difference between V1
and V2 . That voltage drop causes a current through Rgain , and since the feedback loops of the two
input op-amps draw no current, that same amount of current through Rgain must be going through
the two R resistors above and below it. This produces a voltage drop between points 3 and 4
equal to:

V3-4 = (V2 - V1)(1 +

2R )
Rgain

The regular differential amplifier on the right-hand side of the circuit then takes this voltage
drop between points 3 and 4, and amplifies it by a gain of 1 (assuming again that all R resistors
are of equal value). Though this looks like a cumbersome way to build a differential amplifier,
it has the distinct advantages of possessing extremely high input impedances on the V 1 and V2
inputs (because they connect straight into the noninverting inputs of their respective op-amps), and
adjustable gain that can be set by a single resistor. Manipulating the above formula a bit, we have
a general expression for overall voltage gain in the instrumentation amplifier:

AV = (1 + 2R )
Rgain
Though it may not be obvious by looking at the schematic, we can change the differential gain
of the instrumentation amplifier simply by changing the value of one resistor: R gain . Yes, we could
still change the overall gain by changing the values of some of the other resistors, but this would
necessitate balanced resistor value changes for the circuit to remain symmetrical. Please note that

344

CHAPTER 8. OPERATIONAL AMPLIFIERS

the lowest gain possible with the above circuit is obtained with Rgain completely open (infinite
resistance), and that gain value is 1.
REVIEW:
An instrumentation amplifier is a differential op-amp circuit providing high input impedances
with ease of gain adjustment through the variation of a single resistor.

8.11

Differentiator and integrator circuits

By introducing electrical reactance into the feedback loops of op-amp amplifier circuits, we can
cause the output to respond to changes in the input voltage over time. Drawing their names from
their respective calculus functions, the integrator produces a voltage output proportional to the
product (multiplication) of the input voltage and time; and the differentiator (not to be confused
with differential ) produces a voltage output proportional to the input voltages rate of change.
Capacitance can be defined as the measure of a capacitors opposition to changes in voltage.
The greater the capacitance, the more the opposition. Capacitors oppose voltage change by creating
current in the circuit: that is, they either charge or discharge in response to a change in applied
voltage. So, the more capacitance a capacitor has, the greater its charge or discharge current will
be for any given rate of voltage change across it. The equation for this is quite simple:

Changing
DC
voltage

i = C dv
dt
The dv/dt fraction is a calculus expression representing the rate of voltage change over time. If
the DC supply in the above circuit were steadily increased from a voltage of 15 volts to a voltage
of 16 volts over a time span of 1 hour, the current through the capacitor would most likely be very
small, because of the very low rate of voltage change (dv/dt = 1 volt / 3600 seconds). However, if
we steadily increased the DC supply from 15 volts to 16 volts over a shorter time span of 1 second,
the rate of voltage change would be much higher, and thus the charging current would be much
higher (3600 times higher, to be exact). Same amount of change in voltage, but vastly different rates
of change, resulting in vastly different amounts of current in the circuit.
To put some definite numbers to this formula, if the voltage across a 47 F capacitor was changing
at a linear rate of 3 volts per second, the current through the capacitor would be (47 F)(3 V/s)
= 141 A.
We can build an op-amp circuit which measures change in voltage by measuring current through
a capacitor, and outputs a voltage proportional to that current:

8.11.

DIFFERENTIATOR AND INTEGRATOR CIRCUITS

345

Differentiator
C

0V

Vin

0V
0V

Vout

The right-hand side of the capacitor is held to a voltage of 0 volts, due to the virtual ground
effect. Therefore, current through the capacitor is solely due to change in the input voltage. A
steady input voltage wont cause a current through C, but a changing input voltage will.
Capacitor current moves through the feedback resistor, producing a drop across it, which is the
same as the output voltage. A linear, positive rate of input voltage change will result in a steady
negative voltage at the output of the op-amp. Conversely, a linear, negative rate of input voltage
change will result in a steady positive voltage at the output of the op-amp. This polarity inversion
from input to output is due to the fact that the input signal is being sent (essentially) to the inverting
input of the op-amp, so it acts like the inverting amplifier mentioned previously. The faster the rate
of voltage change at the input (either positive or negative), the greater the voltage at the output.
The formula for determining voltage output for the differentiator is as follows:

Vout = -RC

dvin
dt

Applications for this, besides representing the derivative calculus function inside of an analog
computer, include rate-of-change indicators for process instrumentation. One such rate-of-change
signal application might be for monitoring (or controlling) the rate of temperature change in a
furnace, where too high or too low of a temperature rise rate could be detrimental. The DC voltage
produced by the differentiator circuit could be used to drive a comparator, which would signal an
alarm or activate a control if the rate of change exceeded a pre-set level.
In process control, the derivative function is used to make control decisions for maintaining a
process at setpoint, by monitoring the rate of process change over time and taking action to prevent
excessive rates of change, which can lead to an unstable condition. Analog electronic controllers use
variations of this circuitry to perform the derivative function.
On the other hand, there are applications where we need precisely the opposite function, called
integration in calculus. Here, the op-amp circuit would generate an output voltage proportional
to the magnitude and duration that an input voltage signal has deviated from 0 volts. Stated
differently, a constant input signal would generate a certain rate of change in the output voltage:
differentiation in reverse. To do this, all we have to do is swap the capacitor and resistor in the
previous circuit:

346

CHAPTER 8. OPERATIONAL AMPLIFIERS

Integrator

Vin

0V
0V
0V

Vout

As before, the negative feedback of the op-amp ensures that the inverting input will be held at
0 volts (the virtual ground). If the input voltage is exactly 0 volts, there will be no current through
the resistor, therefore no charging of the capacitor, and therefore the output voltage will not change.
We cannot guarantee what voltage will be at the output with respect to ground in this condition,
but we can say that the output voltage will be constant.
However, if we apply a constant, positive voltage to the input, the op-amp output will fall negative
at a linear rate, in an attempt to produce the changing voltage across the capacitor necessary to
maintain the current established by the voltage difference across the resistor. Conversely, a constant,
negative voltage at the input results in a linear, rising (positive) voltage at the output. The output
voltage rate-of-change will be proportional to the value of the input voltage.
The formula for determining voltage output for the integrator is as follows:

Vin
dvout
=dt
RC
or
t

Vout = 0

Vin
dt + c
RC

Where,
c = Output voltage at start time (t=0)
One application for this device would be to keep a running total of radiation exposure, or
dosage, if the input voltage was a proportional signal supplied by an electronic radiation detector.
Nuclear radiation can be just as damaging at low intensities for long periods of time as it is at
high intensities for short periods of time. An integrator circuit would take both the intensity (input
voltage magnitude) and time into account, generating an output voltage representing total radiation
dosage.
Another application would be to integrate a signal representing water flow, producing a signal
representing total quantity of water that has passed by the flowmeter. This application of an
integrator is sometimes called a totalizer in the industrial instrumentation trade.
REVIEW:

8.12.

POSITIVE FEEDBACK

347

A differentiator circuit produces a constant output voltage for a steadily changing input voltage.
An integrator circuit produces a steadily changing output voltage for a constant input voltage.
Both types of devices are easily constructed, using reactive components (usually capacitors
rather than inductors) in the feedback part of the circuit.

8.12

Positive feedback

As weve seen, negative feedback is an incredibly useful principle when applied to operational amplifiers. It is what allows us to create all these practical circuits, being able to precisely set gains,
rates, and other significant parameters with just a few changes of resistor values. Negative feedback
makes all these circuits stable and self-correcting.
The basic principle of negative feedback is that the output tends to drive in a direction that
creates a condition of equilibrium (balance). In an op-amp circuit with no feedback, there is no
corrective mechanism, and the output voltage will saturate with the tiniest amount of differential
voltage applied between the inputs. The result is a comparator:
With negative feedback (the output voltage fed back somehow to the inverting input), the
circuit tends to prevent itself from driving the output to full saturation. Rather, the output voltage
drives only as high or as low as needed to balance the two inputs voltages:

Negative feedback

0V

Vout

+
Vout = Vin

Vin

Whether the output is directly fed back to the inverting (-) input or coupled through a set of
components, the effect is the same: the extremely high differential voltage gain of the op-amp will
be tamed and the circuit will respond according to the dictates of the feedback loop connecting
output to inverting input.
Another type of feedback, namely positive feedback, also finds application in op-amp circuits.
Unlike negative feedback, where the output voltage is fed back to the inverting (-) input, with
positive feedback the output voltage is somehow routed back to the noninverting (+) input. In its
simplest form, we could connect a straight piece of wire from output to noninverting input and see
what happens:

348

CHAPTER 8. OPERATIONAL AMPLIFIERS

Positive feedback
+
Vout

The inverting input remains disconnected from the feedback loop, and is free to receive an
external voltage. Lets see what happens if we ground the inverting input:

+
0V

Vout

With the inverting input grounded (maintained at zero volts), the output voltage will be dictated
by the magnitude and polarity of the voltage at the noninverting input. If that voltage happens
to be positive, the op-amp will drive its output positive as well, feeding that positive voltage back
to the noninverting input, which will result in full positive output saturation. On the other hand,
if the voltage on the noninverting input happens to start out negative, the op-amps output will
drive in the negative direction, feeding back to the noninverting input and resulting in full negative
saturation.
What we have here is a circuit whose output is bistable: stable in one of two states (saturated
positive or saturated negative). Once it has reached one of those saturated states, it will tend to
remain in that state, unchanging. What is necessary to get it to switch states is a voltage placed
upon the inverting (-) input of the same polarity, but of a slightly greater magnitude. For example,
if our circuit is saturated at an output voltage of +12 volts, it will take an input voltage at the
inverting input of at least +12 volts to get the output to change. When it changes, it will saturate
fully negative.
So, an op-amp with positive feedback tends to stay in whatever output state its already in. It
latches between one of two states, saturated positive or saturated negative. Technically, this is
known as hysteresis.
Hysteresis can be a useful property for a comparator circuit to have. As weve seen before,
comparators can be used to produce a square wave from any sort of ramping waveform (sine wave,
triangle wave, sawtooth wave, etc.) input. If the incoming AC waveform is noise-free (that is, a
pure waveform), a simple comparator will work just fine.

8.12.

POSITIVE FEEDBACK

349

+V

Vin

Vout

-V
Square wave
output voltage
DC reference
voltage

AC input
voltage
A "clean" AC input waveform produces predictable
transition points on the output voltage square wave

However, if there exist any anomalies in the waveform such as harmonics or spikes which cause
the voltage to rise and fall significantly within the timespan of a single cycle, a comparators output
might switch states unexpectedly:
+V

Vin

Vout

-V
Square wave
output voltage
DC reference
voltage

AC input
voltage

Any time there is a transition through the reference voltage level, no matter how tiny that
transition may be, the output of the comparator will switch states, producing a square wave with
glitches.
If we add a little positive feedback to the comparator circuit, we will introduce hysteresis into
the output. This hysteresis will cause the output to remain in its current state unless the AC input
voltage undergoes a major change in magnitude.

350

CHAPTER 8. OPERATIONAL AMPLIFIERS

+V

Vin

Vout

-V

Positive feedback
resistor

What this feedback resistor creates is a dual-reference for the comparator circuit. The voltage
applied to the noninverting (+) input as a reference which to compare with the incoming AC
voltage changes depending on the value of the op-amps output voltage. When the op-amp output is
saturated positive, the reference voltage at the noninverting input will be more positive than before.
Conversely, when the op-amp output is saturated negative, the reference voltage at the noninverting
input will be more negative than before. The result is easier to understand on a graph:

square wave
output voltage

DC reference voltages
upper
center
lower

AC input
voltage

When the op-amp output is saturated positive, the upper reference voltage is in effect, and the
output wont drop to a negative saturation level unless the AC input rises above that upper reference
level. Conversely, when the op-amp output is saturated negative, the lower reference voltage is in
effect, and the output wont rise to a positive saturation level unless the AC input drops below that
lower reference level. The result is a clean square-wave output again, despite significant amounts of
distortion in the AC input signal. In order for a glitch to cause the comparator to switch from
one state to another, it would have to be at least as big (tall) as the difference between the upper
and lower reference voltage levels, and at the right point in time to cross both those levels.
Another application of positive feedback in op-amp circuits is in the construction of oscillator
circuits. An oscillator is a device that produces an alternating (AC), or at least pulsing, output
voltage. Technically, it is known as an astable device: having no stable output state (no equilibrium
whatsoever). Oscillators are very useful devices, and they are easily made with just an op-amp and
a few external components.

8.13.

PRACTICAL CONSIDERATIONS

351

Oscillator circuit using positive feedback


C

Vramp

Vout

+
R

R
Vref

Vref
Vramp

Vout is a square wave just like Vref, only taller


When the output is saturated positive, the Vref will be positive, and the capacitor will charge
up in a positive direction. When Vramp exceeds Vref by the tiniest margin, the output will saturate
negative, and the capacitor will charge in the opposite direction (polarity). Oscillation occurs because
the positive feedback is instantaneous and the negative feedback is delayed (by means of an RC time
constant). The frequency of this oscillator may be adjusted by varying the size of any component.
REVIEW:
Negative feedback creates a condition of equilibrium (balance). Positive feedback creates a
condition of hysteresis (the tendency to latch in one of two extreme states).
An oscillator is a device producing an alternating or pulsing output voltage.

8.13

Practical considerations

Real operational have some imperfections compared to an ideal model. A real device deviates from
a perfect difference amplifier. One minus one may not be zero. It may have have an offset like an
analog meter which is not zeroed. The inputs may draw current. The characteristics may drift with
age and temperature. Gain may be reduced at high frequencies, and phase may shift from input to
output. These imperfection may cause no noticable errors in some applications, unacceptable errors
in others. In some cases these errors may be compensated for. Sometimes a higher quality, higher
cost device is required.

352

CHAPTER 8. OPERATIONAL AMPLIFIERS

8.13.1

Common-mode gain

As stated before, an ideal differential amplifier only amplifies the voltage difference between its two
inputs. If the two inputs of a differential amplifier were to be shorted together (thus ensuring zero
potential difference between them), there should be no change in output voltage for any amount of
voltage applied between those two shorted inputs and ground:

Vout

Vout should remain the same


regardless of Vcommon-mode

Vcommon-mode

Voltage that is common between either of the inputs and ground, as V commonmode is in
this case, is called common-mode voltage. As we vary this common voltage, the perfect differential
amplifiers output voltage should hold absolutely steady (no change in output for any arbitrary
change in common-mode input). This translates to a common-mode voltage gain of zero.

AV =

Change in Vout
Change in Vin

. . . if change in Vout = 0 . . .
0
=0
Change in Vin
AV = 0
The operational amplifier, being a differential amplifier with high differential gain, would ideally
have zero common-mode gain as well. In real life, however, this is not easily attained. Thus,
common-mode voltages will invariably have some effect on the op-amps output voltage.
The performance of a real op-amp in this regard is most commonly measured in terms of its
differential voltage gain (how much it amplifies the difference between two input voltages) versus
its common-mode voltage gain (how much it amplifies a common-mode voltage). The ratio of the
former to the latter is called the common-mode rejection ratio, abbreviated as CMRR:

CMRR =

Differential AV
Common-mode AV

An ideal op-amp, with zero common-mode gain would have an infinite CMRR. Real op-amps
have high CMRRs, the ubiquitous 741 having something around 70 dB, which works out to a little
over 3,000 in terms of a ratio.

8.13.

PRACTICAL CONSIDERATIONS

353

Because the common mode rejection ratio in a typical op-amp is so high, common-mode gain
is usually not a great concern in circuits where the op-amp is being used with negative feedback.
If the common-mode input voltage of an amplifier circuit were to suddenly change, thus producing
a corresponding change in the output due to common-mode gain, that change in output would be
quickly corrected as negative feedback and differential gain (being much greater than common-mode
gain) worked to bring the system back to equilibrium. Sure enough, a change might be seen at the
output, but it would be a lot smaller than what you might expect.
A consideration to keep in mind, though, is common-mode gain in differential op-amp circuits
such as instrumentation amplifiers. Outside of the op-amps sealed package and extremely high
differential gain, we may find common-mode gain introduced by an imbalance of resistor values. To
demonstrate this, well run a SPICE analysis on an instrumentation amplifier with inputs shorted
together (no differential voltage), imposing a common-mode voltage to see what happens. First,
well run the analysis showing the output voltage of a perfectly balanced circuit. We should expect
to see no change in output voltage as the common-mode voltage changes:

1 +

V1
0

R3

E1

9
R1

Rgain
Rjump
(jumper
wire)

5
5

R2

R4

E2

instrumentation amplifier
v1 1 0
rin1 1 0 9e12
rjump 1 4 1e-12
rin2 4 0 9e12
e1 3 0 1 2 999k
e2 6 0 4 5 999k
e3 9 0 8 7 999k
rload 9 0 10k
r1 2 3 10k
rgain 2 5 10k
r2 5 6 10k
r3 3 7 10k
r4 7 9 10k
r5 6 8 10k
r6 8 0 10k
.dc v1 0 10 1

8
R5

E3

Vout

R6
8

354

CHAPTER 8. OPERATIONAL AMPLIFIERS

.print dc v(9)
.end
v1
0.000E+00
1.000E+00
2.000E+00
3.000E+00
4.000E+00
5.000E+00
6.000E+00
7.000E+00
8.000E+00
9.000E+00
1.000E+01

v(9)
0.000E+00
1.355E-16
2.710E-16
0.000E+00
5.421E-16
0.000E+00
0.000E+00
0.000E+00
1.084E-15
-1.084E-15
0.000E+00

As you can see, the output voltage v(9)


hardly changes at all for a common-mode
input voltage (v1) that sweeps from 0
to 10 volts.

Aside from very small deviations (actually due to quirks of SPICE rather than real behavior
of the circuit), the output remains stable where it should be: at 0 volts, with zero input voltage
differential. However, lets introduce a resistor imbalance in the circuit, increasing the value of R 5
from 10,000 to 10,500 , and see what happens (the netlist has been omitted for brevity the
only thing altered is the value of R5 ):
v1
0.000E+00
1.000E+00
2.000E+00
3.000E+00
4.000E+00
5.000E+00
6.000E+00
7.000E+00
8.000E+00
9.000E+00
1.000E+01

v(9)
0.000E+00
-2.439E-02
-4.878E-02
-7.317E-02
-9.756E-02
-1.220E-01
-1.463E-01
-1.707E-01
-1.951E-01
-2.195E-01
-2.439E-01

This time we see a significant variation


(from 0 to 0.2439 volts) in output voltage
as the common-mode input voltage sweeps
from 0 to 10 volts as it did before.

Our input voltage differential is still zero volts, yet the output voltage changes significantly as
the common-mode voltage is changed. This is indicative of a common-mode gain, something were
trying to avoid. More than that, its a common-mode gain of our own making, having nothing to
do with imperfections in the op-amps themselves. With a much-tempered differential gain (actually
equal to 3 in this particular circuit) and no negative feedback outside the circuit, this common-mode
gain will go unchecked in an instrument signal application.
There is only one way to correct this common-mode gain, and that is to balance all the resistor
values. When designing an instrumentation amplifier from discrete components (rather than purchasing one in an integrated package), it is wise to provide some means of making fine adjustments
to at least one of the four resistors connected to the final op-amp to be able to trim away any such
common-mode gain. Providing the means to trim the resistor network has additional benefits as

8.13.

PRACTICAL CONSIDERATIONS

355

well. Suppose that all resistor values are exactly as they should be, but a common-mode gain exists
due to an imperfection in one of the op-amps. With the adjustment provision, the resistance could
be trimmed to compensate for this unwanted gain.
One quirk of some op-amp models is that of output latch-up, usually caused by the commonmode input voltage exceeding allowable limits. If the common-mode voltage falls outside of the
manufacturers specified limits, the output may suddenly latch in the high mode (saturate at full
output voltage). In JFET-input operational amplifiers, latch-up may occur if the common-mode
input voltage approaches too closely to the negative power supply rail voltage. On the TL082 opamp, for example, this occurs when the common-mode input voltage comes within about 0.7 volts of
the negative power supply rail voltage. Such a situation may easily occur in a single-supply circuit,
where the negative power supply rail is ground (0 volts), and the input signal is free to swing to 0
volts.
Latch-up may also be triggered by the common-mode input voltage exceeding power supply rail
voltages, negative or positive. As a rule, you should never allow either input voltage to rise above
the positive power supply rail voltage, or sink below the negative power supply rail voltage, even
if the op-amp in question is protected against latch-up (as are the 741 and 1458 op-amp models).
At the very least, the op-amps behavior may become unpredictable. At worst, the kind of latch-up
triggered by input voltages exceeding power supply voltages may be destructive to the op-amp.
While this problem may seem easy to avoid, its possibility is more likely than you might think.
Consider the case of an operational amplifier circuit during power-up. If the circuit receives full
input signal voltage before its own power supply has had time enough to charge the filter capacitors,
the common-mode input voltage may easily exceed the power supply rail voltages for a short time.
If the op-amp receives signal voltage from a circuit supplied by a different power source, and its own
power source fails, the signal voltage(s) may exceed the power supply rail voltages for an indefinite
amount of time!

8.13.2

Offset voltage

Another practical concern for op-amp performance is voltage offset. That is, effect of having the
output voltage something other than zero volts when the two input terminals are shorted together.
Remember that operational amplifiers are differential amplifiers above all: theyre supposed to
amplify the difference in voltage between the two input connections and nothing more. When that
input voltage difference is exactly zero volts, we would (ideally) expect to have exactly zero volts
present on the output. However, in the real world this rarely happens. Even if the op-amp in
question has zero common-mode gain (infinite CMRR), the output voltage may not be at zero when
both inputs are shorted together. This deviation from zero is called offset.

+15 V

+
-15 V

Vout = +14.7 V (saturated +)

356

CHAPTER 8. OPERATIONAL AMPLIFIERS

A perfect op-amp would output exactly zero volts with both its inputs shorted together and
grounded. However, most op-amps off the shelf will drive their outputs to a saturated level, either
negative or positive. In the example shown above, the output voltage is saturated at a value of
positive 14.7 volts, just a bit less than +V (+15 volts) due to the positive saturation limit of this
particular op-amp. Because the offset in this op-amp is driving the output to a completely saturated
point, theres no way of telling how much voltage offset is present at the output. If the +V/-V split
power supply was of a high enough voltage, who knows, maybe the output would be several hundred
volts one way or the other due to the effects of offset!
For this reason, offset voltage is usually expressed in terms of the equivalent amount of input
voltage differential producing this effect. In other words, we imagine that the op-amp is perfect (no
offset whatsoever), and a small voltage is being applied in series with one of the inputs to force the
output voltage one way or the other away from zero. Being that op-amp differential gains are so
high, the figure for input offset voltage doesnt have to be much to account for what we see with
shorted inputs:

+15 V

Vout = +14.7 V (saturated +)

+
-15 V
Input offset voltage
(internal to the real op-amp,
external to this ideal op-amp)

Offset voltage will tend to introduce slight errors in any op-amp circuit. So how do we compensate
for it? Unlike common-mode gain, there are usually provisions made by the manufacturer to trim
the offset of a packaged op-amp. Usually, two extra terminals on the op-amp package are reserved
for connecting an external trim potentiometer. These connection points are labeled offset null
and are used in this general way:

+15 V

Vout

-15 V
Potentiometer adjusted so that
Vout = 0 volts with inputs shorted together
On single op-amps such as the 741 and 3130, the offset null connection points are pins 1 and 5

8.13.

PRACTICAL CONSIDERATIONS

357

on the 8-pin DIP package. Other models of op-amp may have the offset null connections located on
different pins, and/or require a slightly difference configuration of trim potentiometer connection.
Some op-amps dont provide offset null pins at all! Consult the manufacturers specifications for
details.

8.13.3

Bias current

Inputs on an op-amp have extremely high input impedances. That is, the input currents entering or
exiting an op-amps two input signal connections are extremely small. For most purposes of op-amp
circuit analysis, we treat them as though they dont exist at all. We analyze the circuit as though
there was absolutely zero current entering or exiting the input connections.
This idyllic picture, however, is not entirely true. Op-amps, especially those op-amps with bipolar
transistor inputs, have to have some amount of current through their input connections in order for
their internal circuits to be properly biased. These currents, logically, are called bias currents. Under
certain conditions, op-amp bias currents may be problematic. The following circuit illustrates one
of those problem conditions:

+V

Thermocouple

Vout

+
-V

At first glance, we see no apparent problems with this circuit. A thermocouple, generating
a small voltage proportional to temperature (actually, a voltage proportional to the difference in
temperature between the measurement junction and the reference junction formed when the alloy
thermocouple wires connect with the copper wires leading to the op-amp) drives the op-amp either
positive or negative. In other words, this is a kind of comparator circuit, comparing the temperature
between the end thermocouple junction and the reference junction (near the op-amp). The problem
is this: the wire loop formed by the thermocouple does not provide a path for both input bias
currents, because both bias currents are trying to go the same way (either into the op-amp or out
of it).

+V
I?
Thermocouple
I?

Vout

+
-V

This comparator circuit wont work


In order for this circuit to work properly, we must ground one of the input wires, thus providing
a path to (or from) ground for both currents:

358

CHAPTER 8. OPERATIONAL AMPLIFIERS

+V

Thermocouple

Vout

+
I

-V

This comparator circuit will work


Not necessarily an obvious problem, but a very real one!
Another way input bias currents may cause trouble is by dropping unwanted voltages across
circuit resistances. Take this circuit for example:

+V
Voltage drop due
to bias current:
- Rin +
Vin

Ibias

Vout

+
-V

Voltage at (+) op-amp input


will not be exactly equal to Vin
We expect a voltage follower circuit such as the one above to reproduce the input voltage precisely
at the output. But what about the resistance in series with the input voltage source? If there is
any bias current through the noninverting (+) input at all, it will drop some voltage across R in ,
thus making the voltage at the noninverting input unequal to the actual V in value. Bias currents
are usually in the microamp range, so the voltage drop across Rin wont be very much, unless Rin
is very large. One example of an application where the input resistance (R in ) would be very large
is that of pH probe electrodes, where one electrode contains an ion-permeable glass barrier (a very
poor conductor, with millions of of resistance).
If we were actually building an op-amp circuit for pH electrode voltage measurement, wed
probably want to use a FET or MOSFET (IGFET) input op-amp instead of one built with bipolar
transistors (for less input bias current). But even then, what slight bias currents may remain can
cause measurement errors to occur, so we have to find some way to mitigate them through good
design.
One way to do so is based on the assumption that the two input bias currents will be the same.
In reality, they are often close to being the same, the difference between them referred to as the
input offset current. If they are the same, then we should be able to cancel out the effects of input
resistance voltage drop by inserting an equal amount of resistance in series with the other input,
like this:

8.13.

PRACTICAL CONSIDERATIONS

Ibias

Rin(-)

+V
+

- Rin(+)+

Vout

Ibias

Vin

359

-V

With the additional resistance added to the circuit, the output voltage will be closer to V in than
before, even if there is some offset between the two input currents.
For both inverting and noninverting amplifier circuits, the bias current compensating resistor is
placed in series with the noninverting (+) input to compensate for bias current voltage drops in the
divider network:

Noninverting amplifier with


compensating resistor
R1

R2

Rcomp

Vin

Vout

Rcomp = R1 // R2

Inverting amplifier with


compensating resistor
R1

R2

Vin
Rcomp

Rcomp = R1 // R2

Vout

360

CHAPTER 8. OPERATIONAL AMPLIFIERS

In either case, the compensating resistor value is determined by calculating the parallel resistance
value of R1 and R2 . Why is the value equal to the parallel equivalent of R1 and R2 ? When using
the Superposition Theorem to figure how much voltage drop will be produced by the inverting (-)
inputs bias current, we treat the bias current as though it were coming from a current source inside
the op-amp and short-circuit all voltage sources (Vin and Vout ). This gives two parallel paths for
bias current (through R1 and through R2 , both to ground). We want to duplicate the bias currents
effect on the noninverting (+) input, so the resistor value we choose to insert in series with that
input needs to be equal to R1 in parallel with R2 .

A related problem, occasionally experienced by students just learning to build operational amplifier circuits, is caused by a lack of a common ground connection to the power supply. It is imperative
to proper op-amp function that some terminal of the DC power supply be common to the ground
connection of the input signal(s). This provides a complete path for the bias currents, feedback
current(s), and for the load (output) current. Take this circuit illustration, for instance, showing a
properly grounded power supply:

R1

R2

1 k

1 k
+V

null
+

-V

6V

Here, arrows denote the path of electron flow through the power supply batteries, both for
powering the op-amps internal circuitry (the potentiometer inside of it that controls output
voltage), and for powering the feedback loop of resistors R1 and R2 . Suppose, however, that the
ground connection for this split DC power supply were to be removed. The effect of doing this is
profound:

8.13.

PRACTICAL CONSIDERATIONS

361

A power supply ground is essential to circuit operation!


R1

R2

1 k

1 k
+V

broken
connection

null
+

-V

6V

No electrons may flow in or out of the op-amps output terminal, because the pathway to the
power supply is a dead end. Thus, no electrons flow through the ground connection to the left of
R1 , neither through the feedback loop. This effectively renders the op-amp useless: it can neither
sustain current through the feedback loop, nor through a grounded load, since there is no connection
from any point of the power supply to ground.

The bias currents are also stopped, because they rely on a path to the power supply and back
to the input source through ground. The following diagram shows the bias currents (only), as they
go through the input terminals of the op-amp, through the base terminals of the input transistors,
and eventually through the power supply terminal(s) and back to ground.

362

CHAPTER 8. OPERATIONAL AMPLIFIERS

Bias current paths shown, through power supply

+V

Ibias
Ibias
+

-V

6V

Without a ground reference on the power supply, the bias currents will have no complete path
for a circuit, and they will halt. Since bipolar junction transistors are current-controlled devices,
this renders the input stage of the op-amp useless as well, as both input transistors will be forced
into cutoff by the complete lack of base current.
REVIEW:
Op-amp inputs usually conduct very small currents, called bias currents, needed to properly
bias the first transistor amplifier stage internal to the op-amps circuitry. Bias currents are
small (in the microamp range), but large enough to cause problems in some applications.
Bias currents in both inputs must have paths to flow to either one of the power supply rails
or to ground. It is not enough to just have a conductive path from one input to the other.
To cancel any offset voltages caused by bias current flowing through resistances, just add an
equivalent resistance in series with the other op-amp input (called a compensating resistor ).
This corrective measure is based on the assumption that the two input bias currents will be
equal.
Any inequality between bias currents in an op-amp constitutes what is called an input offset
current.
It is essential for proper op-amp operation that there be a ground reference on some terminal
of the power supply, to form complete paths for bias currents, feedback current(s), and load
current.

8.13.4

Drift

Being semiconductor devices, op-amps are subject to slight changes in behavior with changes in operating temperature. Any changes in op-amp performance with temperature fall under the category

8.13.

PRACTICAL CONSIDERATIONS

363

of op-amp drift. Drift parameters can be specified for bias currents, offset voltage, and the like.
Consult the manufacturers data sheet for specifics on any particular op-amp.
To minimize op-amp drift, we can select an op-amp made to have minimum drift, and/or we
can do our best to keep the operating temperature as stable as possible. The latter action may
involve providing some form of temperature control for the inside of the equipment housing the
op-amp(s). This is not as strange as it may first seem. Laboratory-standard precision voltage
reference generators, for example, are sometimes known to employ ovens for keeping their sensitive
components (such as zener diodes) at constant temperatures. If extremely high accuracy is desired
over the usual factors of cost and flexibility, this may be an option worth looking at.
REVIEW:
Op-amps, being semiconductor devices, are susceptible to variations in temperature. Any
variations in amplifier performance resulting from changes in temperature is known as drift.
Drift is best minimized with environmental temperature control.

8.13.5

Frequency response

With their incredibly high differential voltage gains, op-amps are prime candidates for a phenomenon
known as feedback oscillation. Youve probably heard the equivalent audio effect when the volume
(gain) on a public-address or other microphone amplifier system is turned too high: that high pitched
squeal resulting from the sound waveform feeding back through the microphone to be amplified
again. An op-amp circuit can manifest this same effect, with the feedback happening electrically
rather than audibly.
A case example of this is seen in the 3130 op-amp, if it is connected as a voltage follower with the
bare minimum of wiring connections (the two inputs, output, and the power supply connections).
The output of this op-amp will self-oscillate due to its high gain, no matter what the input voltage. To
combat this, a small compensation capacitor must be connected to two specially-provided terminals
on the op-amp. The capacitor provides a high-impedance path for negative feedback to occur within
the op-amps circuitry, thus decreasing the AC gain and inhibiting unwanted oscillations. If the
op-amp is being used to amplify high-frequency signals, this compensation capacitor may not be
needed, but it is absolutely essential for DC or low-frequency AC signal operation.
Some op-amps, such as the model 741, have a compensation capacitor built in to minimize the
need for external components. This improved simplicity is not without a cost: due to that capacitors
presence inside the op-amp, the negative feedback tends to get stronger as the operating frequency
increases (that capacitors reactance decreases with higher frequencies). As a result, the op-amps
differential voltage gain decreases as frequency goes up: it becomes a less effective amplifier at higher
frequencies.
Op-amp manufacturers will publish the frequency response curves for their products. Since a
sufficiently high differential gain is absolutely essential to good feedback operation in op-amp circuits,
the gain/frequency response of an op-amp effectively limits its bandwidth of operation. The circuit
designer must take this into account if good performance is to be maintained over the required range
of signal frequencies.
REVIEW:

364

CHAPTER 8. OPERATIONAL AMPLIFIERS


Due to capacitances within op-amps, their differential voltage gain tends to decrease as the
input frequency increases. Frequency response curves for op-amps are available from the
manufacturer.

8.13.6

Input to output phase shift

In order to illustrate the phase shift from input to output of an operational amplifier (op-amp), the
OPA227 was tested in our lab. The OPA227 was constructed in a typical non-inverting configuration
(Figure 8.1).

Figure 8.1: OPA227 Non-inverting stage


The circuit configuration calls for a signal gain of
=50 dB. The input excitation at
=34 V/V or
Vsrc was set to 10 mVp, and three frequencies of interest: 2.2 kHz, 22 kHz, and 220 MHz. The
OPA227s open loop gain and phase curve vs. frequency is shown in Figure 8.2.

Figure 8.2: AV and vs. Frequency plot


To help predict the closed loop phase shift from input to output, we can use the open loop gain
and phase curve. Since the circuit configuration calls for a closed loop gain, or 1/, of
=50 dB,
the closed loop gain curve intersects the open loop gain curve at approximately 22 kHz. After this

8.14.

OPERATIONAL AMPLIFIER MODELS

365

intersection, the closed loop gain curve rolls off at the typical 20 dB/decade for voltage feedback
amplifiers, and follows the open loop gain curve.
What is actually at work here is the negative feedback from the closed loop modifies the open
loop response. Closing the loop with negative feedback establishes a closed loop pole at 22 kHz.
Much like the dominant pole in the open loop phase curve, we will expect phase shift in the closed
loop response. How much phase shift will we see?
Since the new pole is now at 22 kHz, this is also the -3 dB point as the pole starts to roll off the
closed loop again at 20 dB per decade as stated earlier. As with any pole in basic control theory,
phase shift starts to occur one decade in frequency before the pole, and ends at 90 o of phase shift
one decade in frequency after the pole. So what does this predict for the closed loop response in our
circuit?
This will predict phase shift starting at 2.2 kHz, with 45o of phase shift at the -3 dB point of
22 kHz, and finally ending with 90o of phase shift at 220 kHz. The three Figures shown below are
oscilloscope captures at the frequencies of interest for our OPA227 circuit. Figure 8.3 is set for 2.2
kHz, and no noticeable phase shift is present. Figure 8.4 is set for 220 kHz, and
=45o of phase shift
o

is recorded. Finally, Figure 8.5 is set for 220 MHz, and the expected =90 of phase shift is recorded.
The scope plots were captured using a LeCroy 44x Wavesurfer. The final scope plot used a x1 probe
with the trigger set to HF reject.

Figure 8.3: OPA227 Av=50dB @ 2.2 kHz

8.14

Operational amplifier models

While mention of operational amplifiers typically provokes visions of semiconductor devices built as
integrated circuits on a miniature silicon chip, the first op-amps were actually vacuum tube circuits.
The first commercial, general purpose operational amplifier was manufactured by the George A.

366

CHAPTER 8. OPERATIONAL AMPLIFIERS

Figure 8.4: OPA227 Av=50dB @ 22 kHz

Figure 8.5: OPA227 Av=50dB @ 220 kHz

8.14.

OPERATIONAL AMPLIFIER MODELS

367

Philbrick Researches, Incorporated, in 1952. Designated the K2-W, it was built around two twintriode tubes mounted in an assembly with an octal (8-pin) socket for easy installation and servicing
in electronic equipment chassis of that era. The assembly looked something like this:

The Philbrick Researches


op-amp, model K2-W

approx.
4 inches
GAP/R
MODEL
K2-W

The schematic diagram shows the two tubes, along with ten resistors and two capacitors, a fairly
simple circuit design even by 1952 standards:
The Philbrick Researches op-amp, model K2-W
+300 V

220 k

680 k

510 k

NE-68
12AX7

12AX7

7.5 pF

1 M

Inverting (-)
input

2.2 M
Noninverting (+)
input

Output
221 k
220 k

500 pF

9.1 k

120 k

4.7 M

-300 V

In case youre unfamiliar with the operation of vacuum tubes, they operate similarly to N-channel
depletion-type IGFET transistors: that is, they conduct more current when the control grid (the
dashed line) is made more positive with respect to the cathode (the bent line near the bottom of the
tube symbol), and conduct less current when the control grid is made less positive (or more negative)
than the cathode. The twin triode tube on the left functions as a differential pair, converting the
differential inputs (inverting and noninverting input voltage signals) into a single, amplified voltage
signal which is then fed to the control grid of the left triode of the second triode pair through a

368

CHAPTER 8. OPERATIONAL AMPLIFIERS

voltage divider (1 M 2.2 M). That triode amplifies and inverts the output of the differential
pair for a larger voltage gain, then the amplified signal is coupled to the second triode of the same
dual-triode tube in a noninverting amplifier configuration for a larger current gain. The two neon
glow tubes act as voltage regulators, similar to the behavior of semiconductor zener diodes, to
provide a bias voltage in the coupling between the two single-ended amplifier triodes.
With a dual-supply voltage of +300/-300 volts, this op-amp could only swing its output +/50 volts, which is very poor by todays standards. It had an open-loop voltage gain of 15,000 to
20,000, a slew rate of +/- 12 volts/second, a maximum output current of 1 mA, a quiescent power
consumption of over 3 watts (not including power for the tubes filaments!), and cost about $24 in
1952 dollars. Better performance could have been attained using a more sophisticated circuit design,
but only at the expense of greater power consumption, greater cost, and decreased reliability.
With the advent of solid-state transistors, op-amps with far less quiescent power consumption
and increased reliability became feasible, but many of the other performance parameters remained
about the same. Take for instance Philbricks model P55A, a general-purpose solid-state op-amp
circa 1966. The P55A sported an open-loop gain of 40,000, a slew rate of 1.5 volt/second and an
output swing of +/- 11 volts (at a power supply voltage of +/- 15 volts), a maximum output current
of 2.2 mA, and a cost of $49 (or about $21 for the utility grade version). The P55A, as well as other
op-amps in Philbricks lineup of the time, was of discrete-component construction, its constituent
transistors, resistors, and capacitors housed in a solid brick resembling a large integrated circuit
package.
It isnt very difficult to build a crude operational amplifier using discrete components. A
schematic of one such circuit is shown here:

+V

Q1

Q2
Output

Q3

input (+)

Q5

Q4

Q6

(-) input

A simple operational
amplifier made from
discrete components

-V
While its performance is rather dismal by modern standards, it demonstrates that complexity
is not necessary to create a minimally functional op-amp. Transistors Q 3 and Q4 form the heart of
another differential pair circuit, the semiconductor equivalent of the first triode tube in the K2-W
schematic. As it was in the vacuum tube circuit, the purpose of a differential pair is to amplify and

8.14.

OPERATIONAL AMPLIFIER MODELS

369

convert a differential voltage between the two input terminals to a single-ended output voltage.
With the advent of integrated-circuit (IC) technology, op-amp designs experienced a dramatic
increase in performance, reliability, density, and economy. Between the years of 1964 and 1968, the
Fairchild corporation introduced three models of IC op-amps: the 702, 709, and the still-popular
741. While the 741 is now considered outdated in terms of performance, it is still a favorite among
hobbyists for its simplicity and fault tolerance (short-circuit protection on the output, for instance).
Personal experience abusing many 741 op-amps has led me to the conclusion that it is a hard chip
to kill . . .
The internal schematic diagram for a model 741 op-amp is as follows:
+V

Internal schematic of a model 741 operational amplifier

Q8

Q9

Q12

Q14

Q13

(-) input
Q1

(+) input

R5

Q2

Q15

Q18

R6
Output

Q19
Q3

R10

Q4

R7
Q21

C1
Q16

Q7
Q5

Q20

Q22
Q6

Q11

Q10

Q23

Q17

offset null
offset null
R1

R3

R2

R4

R9

R8

Q24

R11

-V

By integrated circuit standards, the 741 is a very simple device: an example of small-scale
integration, or SSI technology. It would be no small matter to build this circuit using discrete
components, so you can see the advantages of even the most primitive integrated circuit technology
over discrete components where high parts counts are involved.
For the hobbyist, student, or engineer desiring greater performance, there are literally hundreds
of op-amp models to choose from. Many sell for less than a dollar apiece, even retail! Specialpurpose instrumentation and radio-frequency (RF) op-amps may be quite a bit more expensive.
In this section I will showcase several popular and affordable op-amps, comparing and contrasting
their performance specifications. The venerable 741 is included as a benchmark for comparison,
although it is, as I said before, considered an obsolete design.
Listed in Table 8.1 are but a few of the low-cost operational amplifier models widely available
from electronics suppliers. Most of them are available through retail supply stores such as Radio
Shack. All are under $1.00 cost direct from the manufacturer (year 2001 prices). As you can see,
there is substantial variation in performance between some of these units. Take for instance the
parameter of input bias current: the CA3130 wins the prize for lowest, at 0.05 nA (or 50 pA), and
the LM833 has the highest at slightly over 1 A. The model CA3130 achieves its incredibly low bias
current through the use of MOSFET transistors in its input stage. One manufacturer advertises
the 3130s input impedance as 1.5 tera-ohms, or 1.5 x 1012 ! Other op-amps shown here with low

370

CHAPTER 8. OPERATIONAL AMPLIFIERS

Model
number
TL082
LM301A
LM318
LM324
LF353
LF356
LF411
741C
LM833
LM1458
CA3130

Table 8.1: Widely used operational amplifiers


Devices/ Power
Bandwidth Bias
Slew rate
package
supply
current
(count)
(V)
(Mhz)
(nA)
(V/S)
2
12 / 36
4
8
13
1
10 / 36
1
250
0.5
1
10 / 40
15
500
70
4
3 / 32
1
45
0.25
2
12 / 36
4
8
13
1
10 / 36
5
8
12
1
10 / 36
4
20
15
1
10 / 36
1
500
0.5
2
10 / 36
15
1050
7
2
6 / 36
1
800
10
1
5 / 16
15
0.05
10

Output
current
(mA)
17
25
20
20
20
25
25
25
40
45
20

bias current figures use JFET input transistors, while the high bias current models use bipolar input
transistors.
While the 741 is specified in many electronic project schematics and showcased in many textbooks, its performance has long been surpassed by other designs in every measure. Even some
designs originally based on the 741 have been improved over the years to far surpass original design
specifications. One such example is the model 1458, two op-amps in an 8-pin DIP package, which at
one time had the exact same performance specifications as the single 741. In its latest incarnation
it boasts a wider power supply voltage range, a slew rate 50 times as great, and almost twice the
output current capability of a 741, while still retaining the output short-circuit protection feature
of the 741. Op-amps with JFET and MOSFET input transistors far exceed the 741s performance
in terms of bias current, and generally manage to beat the 741 in terms of bandwidth and slew rate
as well.
My own personal recommendations for op-amps are as such: when low bias current is a priority
(such as in low-speed integrator circuits), choose the 3130. For general-purpose DC amplifier work,
the 1458 offers good performance (and you get two op-amps in the space of one package). For an
upgrade in performance, choose the model 353, as it is a pin-compatible replacement for the 1458.
The 353 is designed with JFET input circuitry for very low bias current, and has a bandwidth 4 times
are great as the 1458, although its output current limit is lower (but still short-circuit protected).
It may be more difficult to find on the shelf of your local electronics supply house, but it is just as
reasonably priced as the 1458.
If low power supply voltage is a requirement, I recommend the model 324, as it functions on as
low as 3 volts DC. Its input bias current requirements are also low, and it provides four op-amps
in a single 14-pin chip. Its major weakness is speed, limited to 1 MHz bandwidth and an output
slew rate of only 0.25 volts per s. For high-frequency AC amplifier circuits, the 318 is a very good
general purpose model.
Special-purpose op-amps are available for modest cost which provide better performance specifications. Many of these are tailored for a specific type of performance advantage, such as maximum
bandwidth or minimum bias current. Take for instance the op-amps, both designed for high band-

8.15.

DATA

371

width in Table 8.2.

Model
number
CLC404
CLC425

Table 8.2: High bandwidth operational amplifiers


Devices/ Power
Bandwidth Bias
Slew rate
package
supply
current
(count)
(V)
(Mhz)
(nA)
(V/S)
1
10 / 14
232
44,000
2600
1
5 / 14
1900
40,000
350

Output
current
(mA)
70
90

The CLC404 lists at $21.80 (almost as much as George Philbricks first commercial op-amp,
albeit without correction for inflation), while the CLC425 is quite a bit less expensive at $3.23 per
unit. In both cases high speed is achieved at the expense of high bias currents and restrictive power
supply voltage ranges. Some op-amps, designed for high power output are listed in Table 8.3.

Model
number
LM12CL
LM7171

Table 8.3: High


Devices/ Power
package
supply
(count)
(V)
1
15 / 80
1
5.5 / 36

current operational amplifiers


Bandwidth Bias
Slew rate
current
(Mhz)
(nA)
(V/S)
0.7
1000
9
200
12,000
4100

Output
current
(mA)
13,000
100

Yes, the LM12CL actually has an output current rating of 13 amps (13,000 milliamps)! It lists
at $14.40, which is not a lot of money, considering the raw power of the device. The LM7171, on
the other hand, trades high current output ability for fast voltage output ability (a high slew rate).
It lists at $1.19, about as low as some general purpose op-amps.
Amplifier packages may also be purchased as complete application circuits as opposed to bare
operational amplifiers. The Burr-Brown and Analog Devices corporations, for example, both long
known for their precision amplifier product lines, offer instrumentation amplifiers in pre-designed
packages as well as other specialized amplifier devices. In designs where high precision and repeatability after repair is important, it might be advantageous for the circuit designer to choose such a
pre-engineered amplifier block rather than build the circuit from individual op-amps. Of course,
these units typically cost quite a bit more than individual op-amps.

8.15

Data

Parametrical data for all semiconductor op-amp models except the CA3130 comes from National
Semiconductors online resources, available at this website: (http://www.national.com). Data for
the CA3130 comes from Harris Semiconductors CA3130/CA3130A datasheet (file number 817.4).

Contributors
Contributors to this chapter are listed in chronological order of their contributions, from most recent
to first. See Appendix 2 (Contributor List) for dates and contact information.

372

CHAPTER 8. OPERATIONAL AMPLIFIERS

Wayne Little (June 2007): Author, Input to output phase shift subsection, in Practical
considerations section.

Chapter 9

PRACTICAL ANALOG
SEMICONDUCTOR CIRCUITS
Contents
9.1

Power supply circuits INCOMPLETE . . . . . . . . . . . . . . . . . 373


9.1.1

Unregulated . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374

9.1.2

Linear regulated . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374

9.1.3

Switching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374

9.1.4

Ripple regulated . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 375

9.2

Amplifier circuits PENDING . . . . . . . . . . . . . . . . . . . . . . . 375

9.3

Oscillator circuits INCOMPLETE . . . . . . . . . . . . . . . . . . . . 375


9.3.1

Varactor multiplier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 375

9.4

Phase-locked loops PENDING . . . . . . . . . . . . . . . . . . . . . . 376

9.5

Radio circuits INCOMPLETE . . . . . . . . . . . . . . . . . . . . . . 376

9.6

Computational circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . 379

9.7

Measurement circuits INCOMPLETE . . . . . . . . . . . . . . . . . 400

9.8

Control circuits PENDING . . . . . . . . . . . . . . . . . . . . . . . . 401

*** INCOMPLETE ***

9.1

Power supply circuits INCOMPLETE

There are three major kinds of power supplies: unregulated (also called brute force), linear regulated,
and switching. A fourth type of power supply circuit called the ripple-regulated, is a hybrid between
the brute force and switching designs, and merits a subsection to itself.
373

374

9.1.1

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

Unregulated

An unregulated power supply is the most rudimentary type, consisting of a transformer, rectifier,
and low-pass filter. These power supplies typically exhibit a lot of ripple voltage (i.e. rapidly-varying
instability) and other AC noise superimposed on the DC power. If the input voltage varies, the
output voltage will vary by a proportional amount. The advantage of an unregulated supply is that
its cheap, simple, and efficient.

9.1.2

Linear regulated

A linear regulated supply is simply a brute force (unregulated) power supply followed by a transistor circuit operating in its active, or linear mode, hence the name linear regulator. (Obvious
in retrospect, isnt it?) A typical linear regulator is designed to output a fixed voltage for a wide
range of input voltages, and it simply drops any excess input voltage to allow a maximum output
voltage to the load. This excess voltage drop results in significant power dissipation in the form of
heat. If the input voltage gets too low, the transistor circuit will lose regulation, meaning that it
will fail to keep the voltage steady. It can only drop excess voltage, not make up for a deficiency
in voltage from the brute force section of the circuit. Therefore, you have to keep the input voltage
at least 1 to 3 volts higher than the desired output, depending on the regulator type. This means
the power equivalent of at least 1 to 3 volts multiplied by the full load current will be dissipated
by the regulator circuit, generating a lot of heat. This makes linear regulated power supplies rather
inefficient. Also, to get rid of all that heat they have to use large heat sinks which makes them large,
heavy, and expensive.

9.1.3

Switching

A switching regulated power supply (switcher) is an effort to realize the advantages of both
brute force and linear regulated designs (small, efficient, and cheap, but also clean, stable output
voltage). Switching power supplies work on the principle of rectifying the incoming AC power line
voltage into DC, re-converting it into high-frequency square-wave AC through transistors operated
as on/off switches, stepping that AC voltage up or down by using a lightweight transformer, then
rectifying the transformers AC output into DC and filtering for final output. Voltage regulation
is achieved by altering the duty cycle of the DC-to-AC inversion on the transformers primary
side. In addition to lighter weight because of a smaller transformer core, switchers have another
tremendous advantage over the prior two designs: this type of power supply can be made so totally
independent of the input voltage that it can work on any electric power system in the world; these
are called universal power supplies.
The downside of switchers is that they are more complex, and due to their operation they tend to
generate a lot of high-frequency AC noise on the power line. Most switchers also have significant
ripple voltage on their outputs. With the cheaper types, this noise and ripple can be as bad as for
an unregulated power supply; such low-end switchers arent worthless, because they still provide a
stable average output voltage, and theres the universal input capability.
Expensive switchers are ripple-free and have noise nearly as low as for some a linear types; these
switchers tend to be as expensive as linear supplies. The reason to use an expensive switcher instead
of a good linear is if you need universal power system compatibility or high efficiency. High efficiency,

9.2.

AMPLIFIER CIRCUITS PENDING

375

light weight, and small size are the reasons switching power supplies are almost universally used for
powering digital computer circuitry.

9.1.4

Ripple regulated

A ripple-regulated power supply is an alternative to the linear regulated design scheme: a brute
force power supply (transformer, rectifier, filter) constitutes the front end of the circuit, but a
transistor operated strictly in its on/off (saturation/cutoff) modes transfers DC power to a large
capacitor as needed to maintain the output voltage between a high and a low setpoint. As in
switchers, the transistor in a ripple regulator never passes current while in its active, or linear,
mode for any substantial length of time, meaning that very little energy will be wasted in the form
of heat. However, the biggest drawback to this regulation scheme is the necessary presence of some
ripple voltage on the output, as the DC voltage varies between the two voltage control setpoints.
Also, this ripple voltage varies in frequency depending on load current, which makes final filtering
of the DC power more difficult.
Ripple regulator circuits tend to be quite a bit simpler than switcher circuitry, and they need
not handle the high power line voltages that switcher transistors must handle, making them safer
to work on.

9.2

Amplifier circuits PENDING

9.3

Oscillator circuits INCOMPLETE

9.3.1

Varactor multiplier

A Varactor or variable capacitance diode with a nonlinear capacitance vs frequency characteristic


distorts the applied sinewave f1 in Figure 9.1, generating harmonics, f3.

Vbias
RF blocking
choke
capacitance

fundamental
filter
f1

varicap
diode

harmonic
filter
Resonant
inductor

f3

DC blocking
capacitor

voltage

Figure 9.1: Varactor diode, having a nonlinear capacitance vs voltage characteristic, serves in frequency multiplier.

376

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

The fundamental filter passes f1, blocking the harmonics from returning to the generator. The
choke passes DC, and blocks radio frequencies (RF) from entering the Vbias supply. The harmonic
filter passes the desired harmonic, say the 3rd, to the output, f3. The capacitor at the bottom of the
inductor is a large value, low reactance, to block DC but ground the inductor for RF. The varicap
diode in parallel with the indctor constitutes a parallel resonant network. It is tuned to the desired
harmonic. Note that the reverse bias, Vbias , is fixed.
The varicap multiplier is primarily used to generate microwave signals which cannot be directly
produced by oscillators. The lumped circuit representation in Figure 3.80 is actually stripline or
waveguide sections. Frequenies up to hundreds of gHz may be produced by varactor multipliers.

9.4

Phase-locked loops PENDING

9.5

Radio circuits INCOMPLETE

(b)
L1
240

C1
365
pF

(a)

C2
1000
pf

(c)
(d)

Figure 9.2: (a) Crystal radio. (b) Modulated RF at antenna. (c) Rectified RF at diode cathode,
without C2 filter capacitor. (d) Demodualted audio to headphones.

An antenna ground system, tank circuit, peak detector, and headphones are the the main components of a crystal radio. See Figure 9.2 (a). The antenna absorbs transimtted radio signals (b)
which flow to ground via the other components. The combination of C1 and L1 comprise a resonant
circuit, refered to as a tank circuit. Its purpose is to select one out of many available radios signals.
The variable capacitor C1 allows for tuning to the various signals. The diode passes the positive
half cycles of the RF, removing the negative half cycles (c). C2 is sized to filter the radio frequencies
from the RF envelope (c), passing audio frequencies (d) to the headset. Note that no power supply
is required for a crystal radio. A germanium diode, which has a lower forward voltage drop provides
greater sensitvity than a silicon diode.
The PIN diodes are arranged in a -attenuator network. The anti-series diodes cancel some
harmonic distortion compared with a single series diode. The fixed 1.25 V supply forward biases
the parallel diodes, which not only conducting DC current from ground via the resistors, but also,
conduct RF to ground through the diodes capacitors. The control voltage V control , increases current
through the parallel diodes as it increases. This decreases the resistance and attenuation, passing
more RF from input to output. Attenuation is about 3 dB at Vcontrol = 5 V. Attenuation is 40 dB
at Vcontrol = 1 V with flat frequency response to 2 gHz. At Vcontrol = 0.5 V, attenuation is 80 dB at
10 MHz. However, the frequency response varies too much to use. [?]

9.5.
RADIO CIRCUITS INCOMPLETE

Figure 9.3: Regency TR1: First mass produced transistor radio, 1954.

560

560

268
pF
volume
1K
2 F

20,000 pF
Q2

Q1

Q3

10,000
pF
50,000
pF

1000 pF

50,000
pF

1000 pF

20,000
pF

Q4

1000
pF

3.9 K
1K

10 K
40
F

1000
pF
2.2 K

2.7 K
560
2.2 K

2.7 K

40 F

33 K

2.2 K

470 K

100 K

5 F

22.5 V

377

378

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

10 V

10 V

1K

1K

Transmitter

Receiver
10 V
Transmit
Receive

1K

Figure 9.4: PIN diode T/R switch disconnects receiver from antenna during transmit.

left antenna

right antenna

right

1K

5V

RFC

1K
Receiver

left

-5 V

1K

Figure 9.5: PIN diode antenna switch for direction finder receiver.

Vcontrol = 0 to 5 V
47 nF

47 nF

330

150

330
1.25 V

47 nF

150
330

47 nF

47 nF

5 V less atten
Figure 9.6: PIN diode attenuator: PIN diodes function as voltage variable resistors. After Lin [?].

9.6.

9.6

COMPUTATIONAL CIRCUITS

379

Computational circuits

When someone mentions the word computer, a digital device is what usually comes to mind.
Digital circuits represent numerical quantities in binary format: patterns of 1s and 0s represented
by a multitude of transistor circuits operating in saturated or cutoff states. However, analog circuitry
may also be used to represent numerical quantities and perform mathematical calculations, by using
variable voltage signals instead of discrete on/off states.
Here is a simple example of binary (digital) representation versus analog representation of the
number twenty-five:

A digital circuit representing the number 25:


1
2
4
8
16
32

16 + 8 + 1 = 25

An analog circuit representing the number 25:

Voltmeter
100 V

50

100

380

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

Digital circuits are very different from circuits built on analog principles. Digital computational
circuits can be incredibly complex, and calculations must often be performed in sequential steps to
obtain a final answer, much as a human being would perform arithmetical calculations in steps with
pencil and paper. Analog computational circuits, on the other hand, are quite simple in comparison,
and perform their calculations in continuous, real-time fashion. There is a disadvantage to using
analog circuitry to represent numbers, though: imprecision. The digital circuit shown above is
representing the number twenty-five, precisely. The analog circuit shown above may or may not be
exactly calibrated to 25.000 volts, but is subject to drift and error.
In applications where precision is not critical, analog computational circuits are very practical
and elegant. Shown here are a few op-amp circuits for performing analog computation:

Analog summer (adder) circuit


1 k

1 k

R
Input1
Input2

Output

Output = Input1 + Input2

Analog subtractor circuit


R

Input(-)

Output

+
R

Input(+)

Output = Input(+) - Input(-)

9.6.

COMPUTATIONAL CIRCUITS

381

Analog averager circuit

R
Input2

Output

Input1

(Buffer optional)

Output =

Input1 + Input2
2

Analog inverter (sign reverser) circuit


R

Input

Output

Output = - Input

Analog "multiply-by-constant" circuit


K

Input

Output = (K)(Input)

Output

382

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

Analog "divide-by-constant" circuit

Input

Output
(Buffer optional)

Input
Output =

Analog inverting "multiply/divideby-constant" circuit


K
Input

Output

Output = - (K)(Input)

Each of these circuits may be used in modular fashion to create a circuit capable of multiple
calculations. For instance, suppose that we needed to subtract a certain fraction of one variable
from another variable. By combining a divide-by-constant circuit with a subtractor circuit, we
could obtain the required function:

9.6.

COMPUTATIONAL CIRCUITS

383

Divide-by-constant

Input2
K

Subtractor
R

Output

+
R

Input1
Output = Input1 -

Input2
K

Devices called analog computers used to be common in universities and engineering shops, where
dozens of op-amp circuits could be patched together with removable jumper wires to model mathematical statements, usually for the purpose of simulating some physical process whose underlying
equations were known. Digital computers have made analog computers all but obsolete, but analog computational circuitry cannot be beaten by digital in terms of sheer elegance and economy of
necessary components.
Analog computational circuitry excels at performing the calculus operations integration and
differentiation with respect to time, by using capacitors in an op-amp feedback loop. To fully
understand these circuits operation and applications, though, we must first grasp the meaning of
these fundamental calculus concepts. Fortunately, the application of op-amp circuits to real-world
problems involving calculus serves as an excellent means to teach basic calculus. In the words of
John I. Smith, taken from his outstanding textbook, Modern Operational Circuit Design:
A note of encouragement is offered to certain readers: integral calculus is one of the
mathematical disciplines that operational [amplifier] circuitry exploits and, in the process,
rather demolishes as a barrier to understanding. (pg. 4)
Mr. Smiths sentiments on the pedagogical value of analog circuitry as a learning tool for mathematics are not unique. Consider the opinion of engineer George Fox Lang, in an article he wrote for
the August 2000 issue of the journal Sound and Vibration, entitled, Analog was not a Computer
Trademark!:
Creating a real physical entity (a circuit) governed by a particular set of equations
and interacting with it provides unique insight into those mathematical statements. There
is no better way to develop a gut feel for the interplay between physics and mathematics
than to experience such an interaction. The analog computer was a powerful interdisciplinary teaching tool; its obsolescence is mourned by many educators in a variety of
fields. (pg. 23)

384

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

Differentiation is the first operation typically learned by beginning calculus students. Simply
put, differentiation is determining the instantaneous rate-of-change of one variable as it relates to
another. In analog differentiator circuits, the independent variable is time, and so the rates of change
were dealing with are rates of change for an electronic signal (voltage or current) with respect to
time.
Suppose we were to measure the position of a car, traveling in a direct path (no turns), from its
starting point. Let us call this measurement, x. If the car moves at a rate such that its distance from
start increases steadily over time, its position will plot on a graph as a linear function (straight
line):

x
Position
x

Time
If we were to calculate the derivative of the cars position with respect to time (that is, determine
the rate-of-change of the cars position with respect to time), we would arrive at a quantity representing the cars velocity. The differentiation function is represented by the fractional notation d/d,
so when differentiating position (x ) with respect to time (t), we denote the result (the derivative)
as dx/dt:

x
Position

Velocity
dx
dt

Time

Time

For a linear graph of x over time, the derivate of position (dx/dt), otherwise and more commonly
known as velocity, will be a flat line, unchanging in value. The derivative of a mathematical function
may be graphically understood as its slope when plotted on a graph, and here we can see that the

9.6.

COMPUTATIONAL CIRCUITS

385

position (x ) graph has a constant slope, which means that its derivative (dx/dt) must be constant
over time.
Now, suppose the distance traveled by the car increased exponentially over time: that is, it began
its travel in slow movements, but covered more additional distance with each passing period in time.
We would then see that the derivative of position (dx/dt), otherwise known as velocity (v ), would
not be constant over time, but would increase:

x
Position

Velocity
dx
dt

Time

Time

The height of points on the velocity graph correspond to the rates-of-change, or slope, of points
at corresponding times on the position graph:

Position

Velocity
dx
dt

Time

Time

What does this have to do with analog electronic circuits? Well, if we were to have an analog
voltage signal represent the cars position (think of a huge potentiometer whose wiper was attached
to the car, generating a voltage proportional to the cars position), we could connect a differentiator
circuit to this signal and have the circuit continuously calculate the cars velocity, displaying the
result via a voltmeter connected to the differentiator circuits output:

386

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

Differentiator

+
x V
Position

- dx
V
dt
Velocity
+

x
Recall from the last chapter that a differentiator circuit outputs a voltage proportional to the
input voltages rate-of-change over time (d/dt). Thus, if the input voltage is changing over time at
a constant rate, the output voltage will be at a constant value. If the car moves in such a way that
its elapsed distance over time builds up at a steady rate, then that means the car is traveling at a
constant velocity, and the differentiator circuit will output a constant voltage proportional to that
velocity. If the cars elapsed distance over time changes in a non-steady manner, the differentiator
circuits output will likewise be non-steady, but always at a level representative of the inputs rateof-change over time.
Note that the voltmeter registering velocity (at the output of the differentiator circuit) is connected in reverse polarity to the output of the op-amp. This is because the differentiator circuit
shown is inverting: outputting a negative voltage for a positive input voltage rate-of-change. If
we wish to have the voltmeter register a positive value for velocity, it will have to be connected to
the op-amp as shown. As impractical as it may be to connect a giant potentiometer to a moving
object such as an automobile, the concept should be clear: by electronically performing the calculus
function of differentiation on a signal representing position, we obtain a signal representing velocity.
Beginning calculus students learn symbolic techniques for differentiation. However, this requires
that the equation describing the original graph be known. For example, calculus students learn
how to take a function such as y = 3x and find its derivative with respect to x (d/dx ), 3, simply
by manipulating the equation. We may verify the accuracy of this manipulation by comparing the
graphs of the two functions:

9.6.

COMPUTATIONAL CIRCUITS

387
d 3x = 3
dx

y = 3x

(y = 3)

y
x = 2; y = 3

x = 2.5; slope = 3
x = 2; slope = 3
x = 1; slope = 3
x

x = 1; y = 3

x = 2.5; y = 3
x

Nonlinear functions such as y = 3x2 may also be differentiated by symbolic means. In this case,
the derivative of y = 3x2 with respect to x is 6x :
d 3x2 = 6x
dx

y = 3x2

(y = 6x)

x = 2; slope = 12
y

x = 2; y = 12

x = 1; y = 6
x = 1; slope = 6
x
x = 0; slope = 0

x
x = 0; y = 0

In real life, though, we often cannot describe the behavior of any physical event by a simple
equation like y = 3x, and so symbolic differentiation of the type learned by calculus students may be
impossible to apply to a physical measurement. If someone wished to determine the derivative of our
hypothetical cars position (dx/dt = velocity) by symbolic means, they would first have to obtain
an equation describing the cars position over time, based on position measurements taken from
a real experiment a nearly impossible task unless the car is operated under carefully controlled
conditions leading to a very simple position graph. However, an analog differentiator circuit, by
exploiting the behavior of a capacitor with respect to voltage, current, and time i = C(dv/dt),
naturally differentiates any real signal in relation to time, and would be able to output a signal
corresponding to instantaneous velocity (dx/dt) at any moment. By logging the cars position signal
along with the differentiators output signal using a chart recorder or other data acquisition device,
both graphs would naturally present themselves for inspection and analysis.

388

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

We may take the principle of differentiation one step further by applying it to the velocity signal
using another differentiator circuit. In other words, use it to calculate the rate-of-change of velocity,
which we know is the rate-of-change of position. What practical measure would we arrive at if we
did this? Think of this in terms of the units we use to measure position and velocity. If we were
to measure the cars position from its starting point in miles, then we would probably express its
velocity in units of miles per hour (dx/dt). If we were to differentiate the velocity (measured in miles
per hour) with respect to time, we would end up with a unit of miles per hour per hour. Introductory
physics classes teach students about the behavior of falling objects, measuring position in meters,
velocity in meters per second, and change in velocity over time in meters per second, per second.
This final measure is called acceleration: the rate of change of velocity over time:

x
Position

Velocity
d2x

dx
dt

Time

Acceleration

dt2
Time

Differentiation

Time
Differentiation

The expression d2 x/dt2 is called the second derivative of position (x ) with regard to time (t). If
we were to connect a second differentiator circuit to the output of the first, the last voltmeter would
register acceleration:

9.6.

COMPUTATIONAL CIRCUITS
Differentiator

+
x V
Position

389
Differentiator

- dx
V
dt
Velocity
+

V
Acceleration

+ d2x
2
- dt

Deriving velocity from position, and acceleration from velocity, we see the principle of differentiation very clearly illustrated. These are not the only physical measurements related to each other
in this way, but they are, perhaps, the most common. Another example of calculus in action is the
relationship between liquid flow (q) and liquid volume (v ) accumulated in a vessel over time:

Water
supply

dv = flow
dt

Water

LT v = volume
A Level Transmitter device mounted on a water storage tank provides a signal directly proportional to water level in the tank, which if the tank is of constant cross-sectional area throughout its
height directly equates water volume stored. If we were to take this volume signal and differentiate
it with respect to time (dv/dt), we would obtain a signal proportional to the water flow rate through
the pipe carrying water to the tank. A differentiator circuit connected in such a way as to receive
this volume signal would produce an output signal proportional to flow, possibly substituting for a

390

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

flow-measurement device (Flow Transmitter) installed in the pipe.

Returning to the car experiment, suppose that our hypothetical car were equipped with a tachogenerator on one of the wheels, producing a voltage signal directly proportional to velocity. We
could differentiate the signal to obtain acceleration with one circuit, like this:

Differentiator

v V
Velocity

V
Acceleration

- dv d2x
= 2
dt
dt
+

Gen

By its very nature, the tachogenerator differentiates the cars position with respect to time,
generating a voltage proportional to how rapidly the wheels angular position changes over time.
This provides us with a raw signal already representative of velocity, with only a single step of
differentiation needed to obtain an acceleration signal. A tachogenerator measuring velocity, of
course, is a far more practical example of automobile instrumentation than a giant potentiometer
measuring its physical position, but what we gain in practicality we lose in position measurement.
No matter how many times we differentiate, we can never infer the cars position from a velocity
signal. If the process of differentiation brought us from position to velocity to acceleration, then
somehow we need to perform the reverse process of differentiation to go from velocity to position.
Such a mathematical process does exist, and it is called integration. The integrator circuit may
be used to perform this function of integration with respect to time:

9.6.

COMPUTATIONAL CIRCUITS

391

Integrator

Position

V v dt = x
+

Differentiator

v V
Velocity

- dv d2x
V
= 2
dt
Acceleration
dt
+

Gen

Recall from the last chapter that an integrator circuit outputs a voltage whose rate-of-change
over time is proportional to the input voltages magnitude. Thus, given a constant input voltage,
the output voltage will change at a constant rate. If the car travels at a constant velocity (constant
voltage input to the integrator circuit from the tachogenerator), then its distance traveled will
increase steadily as time progresses, and the integrator will output a steadily changing voltage
proportional to that distance. If the cars velocity is not constant, then neither will the rate-ofchange over time be of the integrator circuits output, but the output voltage will faithfully represent
the amount of distance traveled by the car at any given point in time.
R
The symbol for integration Rlooks something like a very narrow, cursive letter S ( ). The
equation utilizing this symbol ( v dt = x ) tells us that we are integrating velocity (v ) with respect
to time (dt), and obtaining position (x ) as a result.
So, we may express three measures of the cars motion (position, velocity, and acceleration) in
terms of velocity (v ) just as easily as we could in terms of position (x ):

392

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

x
Position
v dt

Velocity
dv
dt

Time

Time
Integration

Acceleration

Time
Differentiation

If we had an accelerometer attached to the car, generating a signal proportional to the rate
of acceleration or deceleration, we could (hypothetically) obtain a velocity signal with one step of
integration, and a position signal with a second step of integration:

9.6.

COMPUTATIONAL CIRCUITS

393

Integrator

+
V a dt = x
Position
-

Integrator

+
a V
Acceleration

Velocity

V a dt = v
+

Accel.

x
Thus, all three measures of the cars motion (position, velocity, and acceleration) may be expressed in terms of acceleration:

x
Position
a dt

Velocity
a dt

Time

Acceleration
a

Time
Integration

Time
Integration

394

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

As you might have suspected, the process of integration may be illustrated in, and applied to,
other physical systems as well. Take for example the water storage tank and flow example shown
earlier. If flow rate is the derivative of tank volume with respect to time (q = dv/dt), then we could
also say that volume is the integral of flow rate with respect to time:

FT f = flow
Water
supply

Water
f dt = volume

If we were to use a Flow Transmitter device to measure water flow, then by time-integration we
could calculate the volume of water accumulated in the tank over time. Although it is theoretically
possible to use a capacitive op-amp integrator circuit to derive a volume signal from a flow signal,
mechanical and digital electronic integrator devices are more suitable for integration over long
periods of time, and find frequent use in the water treatment and distribution industries.
Just as there are symbolic techniques for differentiation, there are also symbolic techniques for
integration, although they tend to be more complex and varied. Applying symbolic integration to
a real-world problem like the acceleration of a car, though, is still contingent on the availability of
an equation precisely describing the measured signal often a difficult or impossible thing to derive
from measured data. However, electronic integrator circuits perform this mathematical function
continuously, in real time, and for any input signal profile, thus providing a powerful tool for scientists
and engineers.
Having said this, there are caveats to the using calculus techniques to derive one type of measurement from another. Differentiation has the undesirable tendency of amplifying noise found
in the measured variable, since the noise will typically appear as frequencies much higher than the
measured variable, and high frequencies by their very nature possess high rates-of-change over time.
To illustrate this problem, suppose we were deriving a measurement of car acceleration from the
velocity signal obtained from a tachogenerator with worn brushes or commutator bars. Points of
poor contact between brush and commutator will produce momentary dips in the tachogenerators
output voltage, and the differentiator circuit connected to it will interpret these dips as very rapid
changes in velocity. For a car moving at constant speed neither accelerating nor decelerating the
acceleration signal should be 0 volts, but noise in the velocity signal caused by a faulty tachogenerator will cause the differentiated (acceleration) signal to contain spikes, falsely indicating brief
periods of high acceleration and deceleration:

9.6.

COMPUTATIONAL CIRCUITS

395

Differentiator

v V
Velocity

- dv
V
dt
Acceleration
+

Gen

Noise voltage present in a signal to be differentiated need not be of significant amplitude to cause
trouble: all that is required is that the noise profile have fast rise or fall times. In other words, any
electrical noise with a high dv/dt component will be problematic when differentiated, even if it is of
low amplitude.

It should be noted that this problem is not an artifact (an idiosyncratic error of the measuring/computing instrument) of the analog circuitry; rather, it is inherent to the process of differentiation. No matter how we might perform the differentiation, noise in the velocity signal will
invariably corrupt the output signal. Of course, if we were differentiating a signal twice, as we did
to obtain both velocity and acceleration from a position signal, the amplified noise signal output by
the first differentiator circuit will be amplified again by the next differentiator, thus compounding
the problem:

396

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS


more noise

little noise

Differentiator

+
x V
Position

even more noise!

Differentiator

- dx
V
dt
Velocity
+

V
Acceleration

+ d2x
2
- dt

Integration does not suffer from this problem, because integrators act as low-pass filters, attenuating high-frequency input signals. In effect, all the high and low peaks resulting from noise on
the signal become averaged together over time, for a diminished net result. One might suppose,
then, that we could avoid all trouble by measuring acceleration directly and integrating that signal
to obtain velocity; in effect, calculating in reverse from the way shown previously:

9.6.

COMPUTATIONAL CIRCUITS

397

Integrator

+
a V
Acceleration

Velocity

V a dt = v
+

Accel.

Unfortunately, following this methodology might lead us into other difficulties, one being a
common artifact of analog integrator circuits known as drift. All op-amps have some amount of
input bias current, and this current will tend to cause a charge to accumulate on the capacitor in
addition to whatever charge accumulates as a result of the input voltage signal. In other words,
all analog integrator circuits suffer from the tendency of having their output voltage drift or
creep even when there is absolutely no voltage input, accumulating error over time as a result.
Also, imperfect capacitors will tend to lose their stored charge over time due to internal resistance,
resulting in drift toward zero output voltage. These problems are artifacts of the analog circuitry,
and may be eliminated through the use of digital computation.

Circuit artifacts notwithstanding, possible errors may result from the integration of one measurement (such as acceleration) to obtain another (such as velocity) simply because of the way
integration works. If the zero calibration point of the raw signal sensor is not perfect, it will output a slight positive or negative signal even in conditions when it should output nothing. Consider
a car with an imperfectly calibrated accelerometer, or one that is influenced by gravity to detect a
slight acceleration unrelated to car motion. Even with a perfect integrating computer, this sensor
error will cause the integrator to accumulate error, resulting in an output signal indicating a change
of velocity when the car is neither accelerating nor decelerating.

398

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

Integrator
(slight positive
voltage)

+
a V
Acceleration

(calibration error)

Velocity

V a dt = v
+

(small rate
of change)

Accel.
(no motion)

As with differentiation, this error will also compound itself if the integrated signal is passed on
to another integrator circuit, since the drifting output of the first integrator will very soon present
a significant positive or negative signal for the next integrator to integrate. Therefore, care should
be taken when integrating sensor signals: if the zero adjustment of the sensor is not perfect, the
integrated result will drift, even if the integrator circuit itself is perfect.

So far, the only integration errors discussed have been artificial in nature: originating from
imperfections in the circuitry and sensors. There also exists a source of error inherent to the process
of integration itself, and that is the unknown constant problem. Beginning calculus students learn
that whenever a function is integrated, there exists an unknown constant (usually represented as
the variable C ) added to the result. This uncertainty is easiest to understand by comparing the
derivatives of several functions differing only by the addition of a constant value:

9.6.

COMPUTATIONAL CIRCUITS

y = 3x2 + 4
y = 3x2

399
d 3x2 + 4 = 6x
dx
d 3x2 = 6x
dx
d 3x2 - 6 = 6x
dx
(y = 6x)

y = 3x2 - 6

Note how each of the parabolic curves (y = 3x2 + C ) share the exact same shape, differing
from each other in regard to their vertical offset. However, they all share the exact same derivative
function: y = (d/dx)( 3x2 + C) = 6x, because they all share identical rates of change (slopes)
at corresponding points along the x axis. While this seems quite natural and expected from the
perspective of differentiation (different equations sharing a common derivative), it usually strikes
beginning students as odd from the perspective of integration, because there are multiple correct
answers for the integral of a function. Going from an equation to its derivative, there is only one
answer, but going from that derivative back to the original equation leads us to a range of correct
solutions. In honor of this uncertainty, the symbolic function of integration is called the indefinite
integral.
When an integrator performs live signal integration with respect to time, the output is the sum
of the integrated input signal over time and an initial value of arbitrary magnitude, representing
the integrators pre-existing output at the time integration began. For example, if I integrate the
velocity of a car driving in a straight line away from a city, calculating
that a constant velocity of
R
50 miles per hour over a time of 2 hours will produce a distance ( v dt) of 100 miles, that does not
necessarily mean the car will be 100 miles away from the city after 2 hours. All it tells us is that
the car will be 100 miles further away from the city after 2 hours of driving. The actual distance
from the city after 2 hours of driving depends on how far the car was from the city when integration
began. If we do not know this initial value for distance, we cannot determine the cars exact distance
from the city after 2 hours of driving.
This same problem appears when we integrate acceleration with respect to time to obtain velocity:

400

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

Integrator

+
a V
Acceleration

V
dt

Velocity

a dt = v + v0
Accel.

Where,
v0 = Initial velocity

x
In this integrator system, the calculated velocity of the car will only be valid if the integrator
circuit is initialized to an output value of zero when the car is stationary (v = 0). Otherwise, the
integrator could very well be outputting a non-zero signal for velocity (v0 ) when the car is stationary,
for the accelerometer cannot tell the difference between a stationary state (0 miles per hour) and
a state of constant velocity (say, 60 miles per hour, unchanging). This uncertainty in integrator
output is inherent to the process of integration, and not an artifact of the circuitry or of the sensor.
In summary, if maximum accuracy is desired for any physical measurement, it is best to measure
that variable directly rather than compute it from other measurements. This is not to say that
computation is worthless. Quite to the contrary, often it is the only practical means of obtaining
a desired measurement. However, the limits of computation must be understood and respected in
order that precise measurements be obtained.

9.7

Measurement circuits INCOMPLETE

Figure 9.7 shows a photodiode amplifier for measuring low levels of light. Best sensitivity and
bandwidth are obtained with a transimpedance amplifier, a current to voltage amplifier, instead
of a conventional operational amplifier. The photodiode remains reverse biased for lowest diode
capacitance, hence wider bandwidth, and lower noise. The feedback resistor sets the gain, the
current to voltage amplification factor. Typical values are 1 to 10 Meg . Higher values yield
higher gain. A capacitor of a few pF may be required to compensate for photodiode capacitance,
and prevents instability at the high gain. The wiring at the summing node must be as compact as
possible. This point is sensitive to circuit board contaminants and must be thoroughly cleaned. The
most sensitive amplifiers contain the photodiode and amplifier within a hybrid microcircuit package
or single die.

9.8.

CONTROL CIRCUITS PENDING

401

Vo
+

Figure 9.7: Photodiode amplifier.

9.8

Control circuits PENDING

Contributors
Contributors to this chapter are listed in chronological order of their contributions, from most recent
to first. See Appendix 2 (Contributor List) for dates and contact information.
Warren Young (August 2002): Initial idea and text for Power supply circuits section. Paragraphs modified by Tony Kuphaldt (changes in vocabulary, plus inclusion of additional concepts).

402

CHAPTER 9. PRACTICAL ANALOG SEMICONDUCTOR CIRCUITS

Chapter 10

ACTIVE FILTERS
Contents

*** PENDING ***

403

404

CHAPTER 10. ACTIVE FILTERS

Chapter 11

DC MOTOR DRIVES
Contents

*** PENDING ***

405

406

CHAPTER 11. DC MOTOR DRIVES

Chapter 12

INVERTERS AND AC MOTOR


DRIVES
Contents

*** PENDING ***

407

408

CHAPTER 12. INVERTERS AND AC MOTOR DRIVES

Chapter 13

ELECTRON TUBES
Contents
13.1
13.2
13.3
13.4
13.5
13.6
13.7
13.8
13.9
13.10
13.11
13.12

13.1

Introduction . . . . . . . . . .
Early tube history . . . . . .
The triode . . . . . . . . . . .
The tetrode . . . . . . . . . .
Beam power tubes . . . . . .
The pentode . . . . . . . . . .
Combination tubes . . . . . .
Tube parameters . . . . . . .
Ionization (gas-filled) tubes .
Display tubes . . . . . . . . .
Microwave tubes . . . . . . .
Tubes versus Semiconductors

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.

409
410
413
415
416
417
418
421
423
427
430
433

Introduction

An often neglected area of study in modern electronics is that of tubes, more precisely known as
vacuum tubes or electron tubes. Almost completely overshadowed by semiconductor, or solid-state
components in most modern applications, tube technology once dominated electronic circuit design.
In fact, the historical transition from electric to electronic circuits really began with tubes,
for it was with tubes that we entered into a whole new realm of circuit function: a way of controlling
the flow of electrons (current) in a circuit by means of another electric signal (in the case of most
tubes, the controlling signal is a small voltage). The semiconductor counterpart to the tube, of
course, is the transistor. Transistors perform much the same function as tubes: controlling the flow
of electrons in a circuit by means of another flow of electrons in the case of the bipolar transistor,
and controlling the flow of electrons by means of a voltage in the case of the field-effect transistor.
In either case, a relatively small electric signal controls a relatively large electric current. This is the
409

410

CHAPTER 13. ELECTRON TUBES

essence of the word electronic, so as to distinguish it from electric, which has more to do with
how electron flow is regulated by Ohms Law and the physical attributes of wire and components.
Though tubes are now obsolete for all but a few specialized applications, they are still a worthy
area of study. If nothing else, it is fascinating to explore the way things used to be done in order
to better appreciate modern technology.

13.2

Early tube history

Thomas Edison, that prolific American inventor, is often credited with the invention of the incandescent lamp. More accurately, it could be said that Edison was the man who perfected the incandescent
lamp. Edisons successful design of 1879 was actually preceded by 77 years by the British scientist
Sir Humphry Davy, who first demonstrated the principle of using electric current to heat a thin strip
of metal (called a filament) to the point of incandescence (glowing white hot).
Edison was able to achieve his success by placing his filament (made of carbonized sewing thread)
inside of a clear glass bulb from which the air had been forcibly removed. In this vacuum, the filament
could glow at white-hot temperatures without being consumed by combustion:

clear, glass bulb


air removed
filament

In the course of his experimentation (sometime around 1883), Edison placed a strip of metal
inside of an evacuated (vacuum) glass bulb along with the filament. Between this metal strip and
one of the filament connections he attached a sensitive ammeter. What he found was that electrons
would flow through the meter whenever the filament was hot, but ceased when the filament cooled
down:

metal strip
A
+

The white-hot filament in Edisons lamp was liberating free electrons into the vacuum of the
lamp, those electrons finding their way to the metal strip, through the galvanometer, and back

13.2.

EARLY TUBE HISTORY

411

to the filament. His curiosity piqued, Edison then connected a fairly high-voltage battery in the
galvanometer circuit to aid the small current:

A
more
current

Sure enough, the presence of the battery created a much larger current from the filament to the
metal strip. However, when the battery was turned around, there was little to no current at all!

no
current ! A

+
-

In effect, what Edison had stumbled upon was a diode! Unfortunately, he saw no practical use
for such a device and proceeded with further refinements in his lamp design.
The one-way electron flow of this device (known as the Edison Effect) remained a curiosity
until J. A. Fleming experimented with its use in 1895. Fleming marketed his device as a valve,
initiating a whole new area of study in electric circuits. Vacuum tube diodes Flemings valves
being no exception are not able to handle large amounts of current, and so Flemings invention
was impractical for any application in AC power, only for small electric signals.
Then in 1906, another inventor by the name of Lee De Forest started playing around with the
Edison Effect, seeing what more could be gained from the phenomenon. In doing so, he made
a startling discovery: by placing a metal screen between the glowing filament and the metal strip
(which by now had taken the form of a plate for greater surface area), the stream of electrons flowing
from filament to plate could be regulated by the application of a small voltage between the metal
screen and the filament:

412

CHAPTER 13. ELECTRON TUBES

The DeForest "Audion" tube

"plate"
-

"grid"
"filament"

control
voltage

plate current can be controlled by the


application of a small control voltage
between the grid and filament!

De Forest called this metal screen between filament and plate a grid. It wasnt just the amount of
voltage between grid and filament that controlled current from filament to plate, it was the polarity
as well. A negative voltage applied to the grid with respect to the filament would tend to choke off
the natural flow of electrons, whereas a positive voltage would tend to enhance the flow. Although
there was some amount of current through the grid, it was very small; much smaller than the current
through the plate.

Perhaps most importantly was his discovery that the small amounts of grid voltage and grid
current were having large effects on the amount of plate voltage (with respect to the filament) and
plate current. In adding the grid to Flemings valve, De Forest had made the valve adjustable: it
now functioned as an amplifying device, whereby a small electrical signal could take control over a
larger electrical quantity.

The closest semiconductor equivalent to the Audion tube, and to all of its more modern tube
equivalents, is an n-channel D-type MOSFET. It is a voltage-controlled device with a large current
gain.

Calling his invention the Audion, he vigorously applied it to the development of communications technology. In 1912 he sold the rights to his Audion tube as a telephone signal amplifier to
the American Telephone and Telegraph Company (AT and T), which made long-distance telephone
communication practical. In the following year he demonstrated the use of an Audion tube for generating radio-frequency AC signals. In 1915 he achieved the remarkable feat of broadcasting voice
signals via radio from Arlington, Virginia to Paris, and in 1916 inaugurated the first radio news
broadcast. Such accomplishments earned De Forest the title Father of Radio in America.

13.3.

13.3

THE TRIODE

413

The triode

De Forests Audion tube came to be known as the triode tube, because it had three elements:
filament, grid, and plate (just as the di in the name diode refers to two elements, filament and
plate). Later developments in diode tube technology led to the refinement of the electron emitter:
instead of using the filament directly as the emissive element, another metal strip called the cathode
could be heated by the filament.
This refinement was necessary in order to avoid some undesired effects of an incandescent filament
as an electron emitter. First, a filament experiences a voltage drop along its length, as current
overcomes the resistance of the filament material and dissipates heat energy. This meant that the
voltage potential between different points along the length of the filament wire and other elements
in the tube would not be constant. For this and similar reasons, alternating current used as a power
source for heating the filament wire would tend to introduce unwanted AC noise in the rest of
the tube circuit. Furthermore, the surface area of a thin filament was limited at best, and limited
surface area on the electron emitting element tends to place a corresponding limit on the tubes
current-carrying capacity.
The cathode was a thin metal cylinder fitting snugly over the twisted wire of the filament. The
cathode cylinder would be heated by the filament wire enough to freely emit electrons, without the
undesirable side effects of actually carrying the heating current as the filament wire had to. The
tube symbol for a triode with an indirectly-heated cathode looks like this:

414

CHAPTER 13. ELECTRON TUBES

plate

grid
cathode

filament

Since the filament is necessary for all but a few types of vacuum tubes, it is often omitted in the
symbol for simplicity, or it may be included in the drawing but with no power connections drawn to
it:

...

...

no filament
shown at all

no connections shown
to filament wires
...

A simple triode circuit is shown to illustrate its basic operation as an amplifier:

Triode amplifier circuit

R
output
voltage

"plate supply"
DC power
source

input
voltage

The low-voltage AC signal connected between the grid and cathode alternately suppresses, then
enhances the electron flow between cathode and plate. This causes a change in voltage on the
output of the circuit (between plate and cathode). The AC voltage and current magnitudes on the
tubes grid are generally quite small compared with the variation of voltage and current in the plate
circuit. Thus, the triode functions as an amplifier of the incoming AC signal (taking high-voltage,

13.4.

THE TETRODE

415

high-current DC power supplied from the large DC source on the right and throttling it by means
of the tubes controlled conductivity).
In the triode, the amount of current from cathode to plate (the controlled current is a function both of grid-to-cathode voltage (the controlling signal) and the plate-to-cathode voltage (the
electromotive force available to push electrons through the vacuum). Unfortunately, neither of these
independent variables have a purely linear effect on the amount of current through the device (often
referred to simply as the plate current). That is, triode current does not necessarily respond in a
direct, proportional manner to the voltages applied.
In this particular amplifier circuit the nonlinearities are compounded, as plate voltage (with
respect to cathode) changes along with the grid voltage (also with respect to cathode) as plate
current is throttled by the tube. The result will be an output voltage waveform that doesnt precisely
resemble the waveform of the input voltage. In other words, the quirkiness of the triode tube and
the dynamics of this particular circuit will distort the waveshape. If we really wanted to get complex
about how we stated this, we could say that the tube introduces harmonics by failing to exactly
reproduce the input waveform.
Another problem with triode behavior is that of stray capacitance. Remember that any time we
have two conductive surfaces separated by an insulating medium, a capacitor will be formed. Any
voltage between those two conductive surfaces will generate an electric field within that insulating
region, potentially storing energy and introducing reactance into a circuit. Such is the case with the
triode, most problematically between the grid and the plate. It is as if there were tiny capacitors
connected between the pairs of elements in the tube:

Cgrid-plate
Cplate-cathode
Cgrid-cathode
Now, this stray capacitance is quite small, and the reactive impedances usually high. Usually,
that is, unless radio frequencies are being dealt with. As we saw with De Forests Audion tube,
radio was probably the prime application for this new technology, so these tiny capacitances
became more than just a potential problem. Another refinement in tube technology was necessary
to overcome the limitations of the triode.

13.4

The tetrode

As the name suggests, the tetrode tube contained four elements: cathode (with the implicit filament,
or heater), grid, plate, and a new element called the screen. Similar in construction to the grid, the
screen was a wire mesh or coil positioned between the grid and plate, connected to a source of positive
DC potential (with respect to the cathode, as usual) equal to a fraction of the plate voltage. When
connected to ground through an external capacitor, the screen had the effect of electrostatically
shielding the grid from the plate. Without the screen, the capacitive linking between the plate and
the grid could cause significant signal feedback at high frequencies, resulting in unwanted oscillations.

416

CHAPTER 13. ELECTRON TUBES

The screen, being of less surface area and lower positive potential than the plate, didnt attract
many of the electrons passing through the grid from the cathode, so the vast majority of electrons
in the tube still flew by the screen to be collected by the plate:

Tetrode amplifier circuit

R
R
"screen"

R
input
voltage

"plate supply"
DC power
source
C

With a constant DC screen voltage, electron flow from cathode to plate became almost exclusively
dependent upon grid voltage, meaning the plate voltage could vary over a wide range with little effect
on plate current. This made for more stable gains in amplifier circuits, and better linearity for more
accurate reproduction of the input signal waveform.
Despite the advantages realized by the addition of a screen, there were some disadvantages as
well. The most significant disadvantage was related to something known as secondary emission.
When electrons from the cathode strike the plate at high velocity, they can cause free electrons to
be jarred loose from atoms in the metal of the plate. These electrons, knocked off the plate by the
impact of the cathode electrons, are said to be secondarily emitted. In a triode tube, secondary
emission is not that great a problem, but in a tetrode with a positively-charged screen grid in close
proximity, these secondary electrons will be attracted to the screen rather than the plate from which
they came, resulting in a loss of plate current. Less plate current means less gain for the amplifier,
which is not good.
Two different strategies were developed to address this problem of the tetrode tube: beam power
tubes and pentodes. Both solutions resulted in new tube designs with approximately the same
electrical characteristics.

13.5

Beam power tubes

In the beam power tube, the basic four-element structure of the tetrode was maintained, but the grid
and screen wires were carefully arranged along with a pair of auxiliary plates to create an interesting
effect: focused beams or sheets of electrons traveling from cathode to plate. These electron beams
formed a stationary cloud of electrons between the screen and plate (called a space charge)
which acted to repel secondary electrons emitted from the plate back to the plate. A set of beamforming plates, each connected to the cathode, were added to help maintain proper electron beam

13.6.

THE PENTODE

417

focus. Grid and screen wire coils were arranged in such a way that each turn or wrap of the screen
fell directly behind a wrap of the grid, which placed the screen wires in the shadow formed by the
grid. This precise alignment enabled the screen to still perform its shielding function with minimal
interference to the passage of electrons from cathode to plate.

grid wires
(cross-sectional view)

beam-forming plates
(2)
"space charge"
- -- - -- -- - - - - ---- -- -- -- - -- -- --- - -

cathode

plate

electron beams

screen wires
(cross-sectional view)
This resulted in lower screen current (and more plate current!) than an ordinary tetrode tube,
with little added expense to the construction of the tube.
Beam power tetrodes were often distinguished from their non-beam counterparts by a different
schematic symbol, showing the beam-forming plates:

The "Beam power" tetrode tube


plate

grid

screen

cathode

13.6

The pentode

Another strategy for addressing the problem of secondary electrons being attracted by the screen
was the addition of a fifth wire element to the tube structure: a suppressor. These five-element
tubes were naturally called pentodes.

418

CHAPTER 13. ELECTRON TUBES

The pentode tube


plate
suppressor

screen

grid

cathode
The suppressor was another wire coil or mesh situated between the screen and the plate, usually
connected directly to ground potential. In some pentode tube designs, the suppressor was internally
connected to the cathode so as to minimize the number of connection pins having to penetrate the
tube envelope:

plate
(suppressor internally
connected to cathode)
screen

grid

cathode
The suppressors job was to repel any secondarily emitted electrons back to the plate: a structural
equivalent of the beam power tubes space charge. This, of course, increased plate current and
decreased screen current, resulting in better gain and overall performance. In some instances it
allowed for greater operating plate voltage as well.

13.7

Combination tubes

Similar in thought to the idea of the integrated circuit, tube designers tried integrating different
tube functions into single tube envelopes to reduce space requirements in more modern tube-type
electronic equipment. A common combination seen within a single glass shell was two either diodes
or two triodes. The idea of fitting pairs of diodes inside a single envelope makes a lot of sense in
light of power supply full-wave rectifier designs, always requiring multiple diodes.
Of course, it would have been quite impossible to combine thousands of tube elements into a
single tube envelope the way that thousands of transistors can be etched onto a single piece of silicon,
but engineers still did their best to push the limits of tube miniaturization and consolidation. Some
of these tubes, whimsically called compactrons, held four or more complete tube elements within a
single envelope.
Sometimes the functions of two different tubes could be integrated into a single, combination
tube in a way that simply worked more elegantly than two tubes ever could. An example of this
was the pentagrid converter, more generally called a heptode, used in some superheterodyne radio
designs. These tubes contained seven elements: 5 grids plus a cathode and a plate. Two of the

13.7.

COMBINATION TUBES

419

grids were normally reserved for signal input, the other three relegated to screening and suppression
(performance-enhancing) functions. Combining the superheterodyne functions of oscillator and signal mixer together in one tube, the signal coupling between these two stages was intrinsic. Rather
than having separate oscillator and mixer circuits, the oscillator creating an AC voltage and the
mixer mixing that voltage with another signal, the pentagrid converters oscillator section created
an electron stream that oscillated in intensity which then directly passed through another grid for
mixing with another signal.
This same tube was sometimes used in a different way: by applying a DC voltage to one of the
control grids, the gain of the tube could be changed for a signal impressed on the other control
grid. This was known as variable-mu operation, because the mu () of the tube (its amplification
factor, measured as a ratio of plate-to-cathode voltage change over grid-to-cathode voltage change
with a constant plate current) could be altered at will by a DC control voltage signal.
Enterprising electronics engineers also discovered ways to exploit such multi-variable capabilities
of lesser tubes such as tetrodes and pentodes. One such way was the so-called ultralinear audio
power amplifier, invented by a pair of engineers named Hafler and Keroes, utilizing a tetrode tube in
combination with a tapped output transformer to provide substantial improvements in amplifier
linearity (decreases in distortion levels). Consider a single-ended triode tube amplifier with an
output transformer coupling power to the speaker:

Speaker

input
voltage

If we substitute a tetrode for a triode in this circuit, we will see improvements in circuit gain resulting from the electrostatic shielding offered by the screen, preventing unwanted feedback between
the plate and control grid:

420

CHAPTER 13. ELECTRON TUBES

Standard
configuration
of tetrode tube
in a single-ended
audio amplifier

Speaker

input
voltage

However, the tetrodes screen may be used for functions other than merely shielding the grid
from the plate. It can also be used as another control element, like the grid itself. If a tap is made
on the transformers primary winding, and this tap connected to the screen, the screen will receive
a voltage that varies with the signal being amplified (feedback). More specifically, the feedback
signal is proportional to the rate-of-change of magnetic flux in the transformer core (d/dt), thus
improving the amplifiers ability to reproduce the input signal waveform at the speaker terminals
and not just in the primary winding of the transformer:

"Ultralinear"
configuration
of tetrode tube
in a single-ended
audio amplifier

Speaker

input
voltage

This signal feedback results in significant improvements in amplifier linearity (and consequently,
distortion), so long as precautions are taken against overpowering the screen with too great a

13.8.

TUBE PARAMETERS

421

positive voltage with respect to the cathode. As a concept, the ultralinear (screen-feedback) design
demonstrates the flexibility of operation granted by multiple grid-elements inside a single tube: a
capability rarely matched by semiconductor components.
Some tube designs combined multiple tube functions in a most economic way: dual plates with
a single cathode, the currents for each of the plates controlled by separate sets of control grids.
Common examples of these tubes were triode-heptode and triode-hexode tubes (a hexode tube is a
tube with four grids, one cathode, and one plate).
Other tube designs simply incorporated separate tube structures inside a single glass envelope
for greater economy. Dual diode (rectifier) tubes were quite common, as were dual triode tubes,
especially when the power dissipation of each tube was relatively low.

Dual triode tube

The 12AX7 and 12AU7 models are common examples of dual-triode tubes, both of low-power
rating. The 12AX7 is especially common as a preamplifier tube in electric guitar amplifier circuits.

13.8

Tube parameters

For bipolar junction transistors, the fundamental measure of amplification is the Beta ratio (),
defined as the ratio of collector current to base current (IC /IB ). Other transistor characteristics
such as junction resistance, which in some amplifier circuits may impact performance as much as ,
are quantified for the benefit of circuit analysis. Electron tubes are no different, their performance
characteristics having been explored and quantified long ago by electrical engineers.
Before we can speak meaningfully on these characteristics, we must define several mathematical
variables used for expressing common voltage, current, and resistance measurements as well as some
of the more complex quantities:

422

CHAPTER 13. ELECTRON TUBES

= amplification factor, pronounced "mu"


(unitless)
gm = mutual conductance, in siemens
Ep = plate-to-cathode voltage
Eg = grid-to-cathode voltage
Ip = plate current
Ik = cathode current
Es = input signal voltage
rp = dynamic plate resistance, in ohms
= delta, the Greek symbol for change

The two most basic measures of an amplifying tubes characteristics are its amplification factor
() and its mutual conductance (gm ), also known as transconductance. Transconductance is defined
here just the same as it is for field-effect transistors, another category of voltage-controlled devices.
Here are the two equations defining each of these performance characteristics:

gm =

Ep
Eg
Ip
Eg

with constant Ip (plate current)

with constant Ep (plate voltage)

Another important, though more abstract, measure of tube performance is its plate resistance.
This is the measurement of plate voltage change over plate current change for a constant value of
grid voltage. In other words, this is an expression of how much the tube acts like a resistor for any
given amount of grid voltage, analogous to the operation of a JFET in its ohmic mode:

rp =

Ep
Ip

with constant Eg (grid voltage)

The astute reader will notice that plate resistance may be determined by dividing the amplification factor by the transconductance:

13.9.

IONIZATION (GAS-FILLED) TUBES

Ep

gm =

Eg

423

Ip
Eg

. . . dividing by gm . . .
Ep
rp =

Eg
Ip
Eg

rp =

rp =

Ep

Eg

Eg

Ip

Ep
Ip

These three performance measures of tubes are subject to change from tube to tube (just as
ratios between two identical bipolar transistors are never precisely the same) and between different
operating conditions. This variability is due partly to the unavoidable nonlinearities of electron tubes
and partly due to how they are defined. Even supposing the existence of a perfectly linear tube,
it will be impossible for all three of these measures to be constant over the allowable ranges of
operation. Consider a tube that perfectly regulates current at any given amount of grid voltage (like
a bipolar transistor with an absolutely constant ): that tubes plate resistance must vary with
plate voltage, because plate current will not change even though plate voltage does.
Nevertheless, tubes were (and are) rated by these values at given operating conditions, and may
have their characteristic curves published just like transistors.

13.9

Ionization (gas-filled) tubes

So far, weve explored tubes which are totally evacuated of all gas and vapor inside their glass
envelopes, properly known as vacuum tubes. With the addition of certain gases or vapors, however,
tubes take on significantly different characteristics, and are able to fulfill certain special roles in
electronic circuits.
When a high enough voltage is applied across a distance occupied by a gas or vapor, or when
that gas or vapor is heated sufficiently, the electrons of those gas molecules will be stripped away
from their respective nuclei, creating a condition of ionization. Having freed the electrons from their
electrostatic bonds to the atoms nuclei, they are free to migrate in the form of a current, making
the ionized gas a relatively good conductor of electricity. In this state, the gas is more properly
referred to as a plasma.
Ionized gas is not a perfect conductor. As such, the flow of electrons through ionized gas will
tend to dissipate energy in the form of heat, thereby helping to keep the gas in a state of ionization.
The result of this is a tube that will begin to conduct under certain conditions, then tend to stay

424

CHAPTER 13. ELECTRON TUBES

in a state of conduction until the applied voltage across the gas and/or the heat-generating current
drops to a minimum level.
The astute observer will note that this is precisely the kind of behavior exhibited by a class of
semiconductor devices called thyristors, which tend to stay on once turned on and tend to stay
off once turned off. Gas-filled tubes, it can be said, manifest this same property of hysteresis.
Unlike their vacuum counterparts, ionization tubes were often manufactured with no filament
(heater) at all. These were called cold-cathode tubes, with the heated versions designated as hotcathode tubes. Whether or not the tube contained a source of heat obviously impacted the characteristics of a gas-filled tube, but not to the extent that lack of heat would impact the performance
of a hard-vacuum tube.
The simplest type of ionization device is not necessarily a tube at all; rather, it is constructed
of two electrodes separated by a gas-filled gap. Simply called a spark gap, the gap between the
electrodes may be occupied by ambient air, other times a special gas, in which case the device must
have a sealed envelope of some kind.

Spark gap
enclosure (optional)

electrodes
A prime application for spark gaps is in overvoltage protection. Engineered not to ionize, or
break down (begin conducting), with normal system voltage applied across the electrodes, the
spark gaps function is to conduct in the event of a significant increase in voltage. Once conducting,
it will act as a heavy load, holding the system voltage down through its large current draw and
subsequent voltage drop along conductors and other series impedances. In a properly engineered
system, the spark gap will stop conducting (extinguish) when the system voltage decreases to a
normal level, well below the voltage required to initiate conduction.
One major caveat of spark gaps is their significantly finite life. The discharge generated by such
a device can be quite violent, and as such will tend to deteriorate the surfaces of the electrodes
through pitting and/or melting.
Spark gaps can be made to conduct on command by placing a third electrode (usually with a
sharp edge or point) between the other two and applying a high voltage pulse between that electrode
and one of the other electrodes. The pulse will create a small spark between the two electrodes,
ionizing part of the pathway between the two large electrodes, and enabling conduction between
them if the applied voltage is high enough:

13.9.

IONIZATION (GAS-FILLED) TUBES

425

Triggered spark gap

main
voltage
source

(high voltage,
high current)

Load

spark gap

third electrode

triggering voltage source


(high voltage, low current)

Spark gaps of both the triggered and untriggered variety can be built to handle huge amounts
of current, some even into the range of mega-amps (millions of amps)! Physical size is the primary
limiting factor to the amount of current a spark gap can safely and reliably handle.

When the two main electrodes are placed in a sealed tube filled with a special gas, a discharge
tube is formed. The most common type of discharge tube is the neon light, used popularly as a
source of colorful illumination, the color of the light emitted being dependent on the type of gas
filling the tube.

Construction of neon lamps closely resembles that of spark gaps, but the operational characteristics are quite different:

426

CHAPTER 13. ELECTRON TUBES


high voltage power supply (AC or DC)

NEON LAMP

electrode

current through the tube


causes the neon gas to glow

electrode

glass tube

small neon
indicator lamp
Neon lamp schematic symbol

By controlling the spacing of the electrodes and the type of gas in the tube, neon lights can
be made to conduct without drawing the excessive currents that spark gaps do. They still exhibit
hysteresis in that it takes a higher voltage to initiate conduction than it does to make them extinguish, and their resistance is definitely nonlinear (the more voltage applied across the tube, the
more current, thus more heat, thus lower resistance). Given this nonlinear tendency, the voltage
across a neon tube must not be allowed to exceed a certain limit, lest the tube be damaged by
excessive temperatures.
This nonlinear tendency gives the neon tube an application other than colorful illumination: it
can act somewhat like a zener diode, clamping the voltage across it by drawing more and more
current if the voltage decreases. When used in this fashion, the tube is known as a glow tube, or
voltage-regulator tube, and was a popular means of voltage regulation in the days of electron tube
circuit design.
Rseries

glow-discharge
voltage regulator
tube

Rload

voltage across load


held relative constant
with variations of voltage
source and load resistance

Please take note of the black dot found in the tube symbol shown above (and in the neon lamp
symbol shown before that). That marker indicates the tube is gas-filled. It is a common marker
used in all gas-filled tube symbols.
One example of a glow tube designed for voltage regulation was the VR-150, with a nominal
regulating voltage of 150 volts. Its resistance throughout the allowable limits of current could vary

13.10.

DISPLAY TUBES

427

from 5 k to 30 k, a 6:1 span. Like zener diode regulator circuits of today, glow tube regulators
could be coupled to amplifying tubes for better voltage regulation and higher load current ranges.
If a regular triode was filled with gas instead of a hard vacuum, it would manifest all the
hysteresis and nonlinearity of other gas tubes with one major advantage: the amount of voltage
applied between grid and cathode would determine the minimum plate-to cathode voltage necessary
to initiate conduction. In essence, this tube was the equivalent of the semiconductor SCR (SiliconControlled Rectifier), and was called the thyratron.
(Simple) Thyratron control circuit
+
Rload
Thyratron
Tube

high voltage
AC source

control
voltage

It should be noted that the schematic shown above is greatly simplified for most purposes and
thyratron tube designs. Some thyratrons, for instance, required that the grid voltage switch polarity
between their on and off states in order to properly work. Also, some thyratrons had more than
one grid!
Thyratrons found use in much the same way as SCRs find use today: controlling rectified AC
to large loads such as motors. Thyratron tubes have been manufactured with different types of gas
fillings for different characteristics: inert (chemically non-reactive) gas, hydrogen gas, and mercury
(vaporized into a gas form when activated). Deuterium, a rare isotope of hydrogen, was used in
some special applications requiring the switching of high voltages.

13.10

Display tubes

In addition to performing tasks of amplification and switching, tubes can be designed to serve as
display devices.
Perhaps the best-known display tube is the cathode ray tube, or CRT. Originally invented as an
instrument to study the behavior of cathode rays (electrons) in a vacuum, these tubes developed
into instruments useful in detecting voltage, then later as video projection devices with the advent
of television. The main difference between CRTs used in oscilloscopes and CRTs used in televisions
is that the oscilloscope variety exclusively use electrostatic (plate) deflection, while televisions use
electromagnetic (coil) deflection. Plates function much better than coils over a wider range of signal
frequencies, which is great for oscilloscopes but irrelevant for televisions, since a television electron
beam sweeps vertically and horizontally at fixed frequencies. Electromagnetic deflection coils are
much preferred in television CRT construction because they do not have to penetrate the glass

428

CHAPTER 13. ELECTRON TUBES

envelope of the tube, thus decreasing the production costs and increasing tube reliability.
An interesting cousin to the CRT is the Cat-Eye or Magic-Eye indicator tube. Essentially,
this tube is a voltage-measuring device with a display resembling a glowing green ring. Electrons
emitted by the cathode of this tube impinge on a fluorescent screen, causing the green-colored light
to be emitted. The shape of the glow produced by the fluorescent screen varies as the amount of
voltage applied to a grid changes:

"Cat-Eye" indicator tube displays

large shadow

slight shadow

minimal shadow

The width of the shadow is directly determined by the potential difference between the control
electrode and the fluorescent screen. The control electrode is a narrow rod placed between the
cathode and the fluorescent screen. If that control electrode (rod) is significantly more negative
than the fluorescent screen, it will deflect some electrons away from the that area of the screen. The
area of the screen shadowed by the control electrode will appear darker when there is a significant
voltage difference between the two. When the control electrode and fluorescent screen are at equal
potential (zero voltage between them), the shadowing effect will be minimal and the screen will be
equally illuminated.
The schematic symbol for a cat-eye tube looks something like this:

"Cat-Eye" or "Magic-Eye"
indicator tube
plate

fluorescent
screen
control
electrode

amplifier
grid
cathode
Here is a photograph of a cat-eye tube, showing the circular display region as well as the glass
envelope, socket (black, at far end of tube), and some of its internal structure:

13.10.

DISPLAY TUBES

429

Normally, only the end of the tube would protrude from a hole in an instrument panel, so the
user could view the circular, fluorescent screen.
In its simplest usage, a cat-eye tube could be operated without the use of the amplifier grid.
However, in order to make it more sensitive, the amplifier grid is used, and it is used like this:

"Cat-Eye" indicator tube circuit

signal

As the signal voltage increases, current through


the tube is choked off. This decreases the voltage
between the plate and the fluorescent screen,
lessening the shadow effect (shadow narrows).
The cathode, amplifier grid, and plate act as a triode to create large changes in plate-to-cathode
voltage for small changes in grid-to-cathode voltage. Because the control electrode is internally
connected to the plate, it is electrically common to it and therefore possesses the same amount of
voltage with respect to the cathode that the plate does. Thus, the large voltage changes induced
on the plate due to small voltage changes on the amplifier grid end up causing large changes in the
width of the shadow seen by whoever is viewing the tube.

430

CHAPTER 13. ELECTRON TUBES

Control electrode negative with


respect to the fluorescent screen.
This is caused by a positive
amplifier grid voltage (with
respect to the cathode).

No voltage between control


electrode and flourescent screen.
This is caused by a negative
amplifier grid voltage (with
respect to the cathode).

Cat-eye tubes were never accurate enough to be equipped with a graduated scale as is the case
with CRTs and electromechanical meter movements, but they served well as null detectors in bridge
circuits, and as signal strength indicators in radio tuning circuits. An unfortunate limitation to the
cat-eye tube as a null detector was the fact that it was not directly capable of voltage indication
in both polarities.

13.11

Microwave tubes

For extremely high-frequency applications (above 1 GHz), the interelectrode capacitances and transittime delays of standard electron tube construction become prohibitive. However, there seems to be
no end to the creative ways in which tubes may be constructed, and several high-frequency electron
tube designs have been made to overcome these challenges.
It was discovered in 1939 that a toroidal cavity made of conductive material called a cavity
resonator surrounding an electron beam of oscillating intensity could extract power from the beam
without actually intercepting the beam itself. The oscillating electric and magnetic fields associated
with the beam echoed inside the cavity, in a manner similar to the sounds of traveling automobiles
echoing in a roadside canyon, allowing radio-frequency energy to be transferred from the beam to a
waveguide or coaxial cable connected to the resonator with a coupling loop. The tube was called an
inductive output tube, or IOT :

The inductive output tube (IOT)


coaxial
output
cable

RF power
output

RF
signal
input
electron beam
toroidal
cavity

DC supply

13.11.

MICROWAVE TUBES

431

Two of the researchers instrumental in the initial development of the IOT, a pair of brothers
named Sigurd and Russell Varian, added a second cavity resonator for signal input to the inductive
output tube. This input resonator acted as a pair of inductive grids to alternately bunch and
release packets of electrons down the drift space of the tube, so the electron beam would be composed
of electrons traveling at different velocities. This velocity modulation of the beam translated into
the same sort of amplitude variation at the output resonator, where energy was extracted from the
beam. The Varian brothers called their invention a klystron.

The klystron tube


coaxial
output
cable

RF
signal
input

RF power
output

Beam
control
electron beam

DC supply
Another invention of the Varian brothers was the reflex klystron tube. In this tube, electrons
emitted from the heated cathode travel through the cavity grids toward the repeller plate, then are
repelled and returned back the way they came (hence the name reflex ) through the cavity grids.
Self-sustaining oscillations would develop in this tube, the frequency of which could be changed by
adjusting the repeller voltage. Hence, this tube operated as a voltage-controlled oscillator.

The reflex klystron tube


cavity
grids

repeller

RF output
cavity
control grid
cathode
As a voltage-controlled oscillator, reflex klystron tubes served commonly as local oscillators

432

CHAPTER 13. ELECTRON TUBES

for radar equipment and microwave receivers:

Reflex klystron tube used as


a voltage-controlled oscillator

Initially developed as low-power devices whose output required further amplification for radio
transmitter use, reflex klystron design was refined to the point where the tubes could serve as power
devices in their own right. Reflex klystrons have since been superseded by semiconductor devices in
the application of local oscillators, but amplification klystrons continue to find use in high-power,
high-frequency radio transmitters and in scientific research applications.

One microwave tube performs its task so well and so cost-effectively that it continues to reign
supreme in the competitive realm of consumer electronics: the magnetron tube. This device forms
the heart of every microwave oven, generating several hundred watts of microwave RF energy used
to heat food and beverages, and doing so under the most grueling conditions for a tube: powered
on and off at random times and for random durations.

Magnetron tubes are representative of an entirely different kind of tube than the IOT and
klystron. Whereas the latter tubes use a linear electron beam, the magnetron directs its electron
beam in a circular pattern by means of a strong magnetic field:

13.12.

TUBES VERSUS SEMICONDUCTORS

433

The magnetron tube


cavity
resonators

electron
cathode

RF output

beam

Once again, cavity resonators are used as microwave-frequency tank circuits, extracting energy
from the passing electron beam inductively. Like all microwave-frequency devices using a cavity
resonator, at least one of the resonator cavities is tapped with a coupling loop: a loop of wire
magnetically coupling the coaxial cable to the resonant structure of the cavity, allowing RF power
to be directed out of the tube to a load. In the case of the microwave oven, the output power is
directed through a waveguide to the food or drink to be heated, the water molecules within acting
as tiny load resistors, dissipating the electrical energy in the form of heat.
The magnet required for magnetron operation is not shown in the diagram. Magnetic flux runs
perpendicular to the plane of the circular electron path. In other words, from the view of the tube
shown in the diagram, you are looking straight at one of the magnetic poles.

13.12

Tubes versus Semiconductors

Devoting a whole chapter in a modern electronics text to the design and function of electron tubes
may seem a bit strange, seeing as how semiconductor technology has all but obsoleted tubes in almost
every application. However, there is merit in exploring tubes not just for historical purposes, but
also for those niche applications that necessitate the qualifying phrase almost every application
in regard to semiconductor supremacy.
In some applications, electron tubes not only continue to see practical use, but perform their
respective tasks better than any solid-state device yet invented. In some cases the performance and
reliability of electron tube technology is far superior.
In the fields of high-power, high-speed circuit switching, specialized tubes such as hydrogen
thyratrons and krytrons are able to switch far larger amounts of current, far faster than any semiconductor device designed to date. The thermal and temporal limits of semiconductor physics place
limitations on switching ability that tubes which do not operate on the same principles are
exempt from.
In high-power microwave transmitter applications, the excellent thermal tolerance of tubes alone
secures their dominance over semiconductors. Electron conduction through semiconducting materials is greatly impacted by temperature. Electron conduction through a vacuum is not. As a

434

CHAPTER 13. ELECTRON TUBES

consequence, the practical thermal limits of semiconductor devices are rather low compared to that
of tubes. Being able to operate tubes at far greater temperatures than equivalent semiconductor
devices allows tubes to dissipate more thermal energy for a given amount of dissipation area, which
makes them smaller and lighter in continuous high power applications.
Another decided advantage of tubes over semiconductor components in high-power applications
is their rebuildability. When a large tube fails, it may be disassembled and repaired at far lower
cost than the purchase price of a new tube. When a semiconductor component fails, large or small,
there is generally no means of repair.
The following photograph shows the front panel of a 1960s vintage 5 kW AM radio transmitter.
One of two Eimac brand power tubes can be seen in a recessed area, behind the glass door.
According to the station engineer who gave the facility tour, the rebuild cost for such a tube is only
$800: quite inexpensive compared to the cost of a new tube, and still quite reasonable in contrast
to the price of a new, comparable semiconductor component!

Tubes, being less complex in their manufacture than semiconductor components, are potentially
cheaper to produce as well, although the huge volume of semiconductor device production in the
world greatly offsets this theoretical advantage. Semiconductor manufacture is quite complex, involving many dangerous chemical substances and necessitating super-clean assembly environments.
Tubes are essentially nothing more than glass and metal, with a vacuum seal. Physical tolerances
are loose enough to permit hand-assembly of vacuum tubes, and the assembly work need not be
done in a clean room environment as is necessary for semiconductor manufacture.
One modern area where electron tubes enjoy supremacy over semiconductor components is in
the professional and high-end audio amplifier markets, although this is partially due to musical culture. Many professional guitar players, for example, prefer tube amplifiers over transistor amplifiers
because of the specific distortion produced by tube circuits. An electric guitar amplifier is designed
to produce distortion rather than avoid distortion as is the case with audio-reproduction amplifiers
(this is why an electric guitar sounds so much different than an acoustical guitar), and the type of
distortion produced by an amplifier is as much a matter of personal taste as it is technical measurement. Since rock music in particular was born with guitarists playing tube-amplifier equipment,
there is a significant level of tube appeal inherent to the genre itself, and this appeal shows itself
in the continuing demand for tubed guitar amplifiers among rock guitarists.

13.12.

TUBES VERSUS SEMICONDUCTORS

435

As an illustration of the attitude among some guitarists, consider the following quote taken from
the technical glossary page of a tube-amplifier website which will remain nameless:
Solid State: A component that has been specifically designed to make a guitar amplifier sound bad. Compared to tubes, these devices can have a very long lifespan, which
guarantees that your amplifier will retain its thin, lifeless, and buzzy sound for a long
time to come.
In the area of audio reproduction amplifiers (music studio amplifiers and home entertainment
amplifiers), it is best for an amplifier to reproduce the musical signal with as little distortion as
possible. Paradoxically, in contrast to the guitar amplifier market where distortion is a design
goal, high-end audio is another area where tube amplifiers enjoy continuing consumer demand.
Though one might suppose the objective, technical requirement of low distortion would eliminate
any subjective bias on the part of audiophiles, one would be very wrong. The market for high-end
tubed amplifier equipment is quite volatile, changing rapidly with trends and fads, driven by
highly subjective claims of magical sound from audio system reviewers and salespeople. As in the
electric guitar world, there is no small measure of cult-like devotion to tube amplifiers among some
quarters of the audiophile world. As an example of this irrationality, consider the design of many
ultra-high-end amplifiers, with chassis built to display the working tubes openly, even though this
physical exposure of the tubes obviously enhances the undesirable effect of microphonics (changes
in tube performance as a result of sound waves vibrating the tube structure).
Having said this, though, there is a wealth of technical literature contrasting tubes against
semiconductors for audio power amplifier use, especially in the area of distortion analysis. More
than a few competent electrical engineers prefer tube amplifier designs over transistors, and are able
to produce experimental evidence in support of their choice. The primary difficulty in quantifying
audio system performance is the uncertain response of human hearing. All amplifiers distort their
input signal to some degree, especially when overloaded, so the question is which type of amplifier
design distorts the least. However, since human hearing is very nonlinear, people do not interpret
all types of acoustic distortion equally, and so some amplifiers will sound better than others even
if a quantitative distortion analysis with electronic instruments indicates similar distortion levels.
To determine what type of audio amplifier will distort a musical signal the least, we must regard
the human ear and brain as part of the whole acoustical system. Since no complete model yet exists
for human auditory response, objective assessment is difficult at best. However, some research
indicates that the characteristic distortion of tube amplifier circuits (especially when overloaded) is
less objectionable than distortion produced by transistors.
Tubes also possess the distinct advantage of low drift over a wide range of operating conditions.
Unlike semiconductor components, whose barrier voltages, ratios, bulk resistances, and junction
capacitances may change substantially with changes in device temperature and/or other operating
conditions, the fundamental characteristics of a vacuum tube remain nearly constant over a wide
range in operating conditions, because those characteristics are determined primarily by the physical
dimensions of the tubes structural elements (cathode, grid(s), and plate) rather than the interactions
of subatomic particles in a crystalline lattice.
This is one of the major reasons solid-state amplifier designers typically engineer their circuits
to maximize power-efficiency even when it compromises distortion performance, because a powerinefficient amplifier dissipates a lot of energy in the form of waste heat, and transistor characteristics
tend to change substantially with temperature. Temperature-induced drift makes it difficult to

436

CHAPTER 13. ELECTRON TUBES

stabilize Q points and other important performance-related measures in an amplifier circuit. Unfortunately, power efficiency and low distortion seem to be mutually exclusive design goals.
For example, class A audio amplifier circuits typically exhibit very low distortion levels, but are
very wasteful of power, meaning that it would be difficult to engineer a solid-state class A amplifier
of any substantial power rating due to the consequent drift of transistor characteristics. Thus, most
solid-state audio amplifier designers choose class B circuit configurations for greater efficiency, even
though class B designs are notorious for producing a type of distortion known as crossover distortion.
However, with tubes it is easy to design a stable class A audio amplifier circuit because tubes are
not as adversely affected by the changes in temperature experienced in a such a power-inefficient
circuit configuration.
Tube performance parameters, though, tend to drift more than semiconductor devices when
measured over long periods of time (years). One major mechanism of tube aging appears to
be vacuum leaks: when air enters the inside of a vacuum tube, its electrical characteristics become
irreversibly altered. This same phenomenon is a major cause of tube mortality, or why tubes typically
do not last as long as their respective solid-state counterparts. When tube vacuum is maintained
at a high level, though, excellent performance and life is possible. An example of this is a klystron
tube (used to produce the high-frequency radio waves used in a radar system) that lasted for 240,000
hours of operation (cited by Robert S. Symons of Litton Electron Devices Division in his informative
paper, Tubes: Still vital after all these years, printed in the April 1998 issue of IEEE Spectrum
magazine).
If nothing else, the tension between audiophiles over tubes versus semiconductors has spurred
a remarkable degree of experimentation and technical innovation, serving as an excellent resource
for those wishing to educate themselves on amplifier theory. Taking a wider view, the versatility
of electron tube technology (different physical configurations, multiple control grids) hints at the
potential for circuit designs of far greater variety than is possible using semiconductors. For this
and other reasons, electron tubes will never be obsolete, but will continue to serve in niche roles,
and to foster innovation for those electronics engineers, inventors, and hobbyists who are unwilling
to let their minds by stifled by convention.

Appendix A-1

ABOUT THIS BOOK

A-1.1

Purpose

They say that necessity is the mother of invention. At least in the case of this book, that adage
is true. As an industrial electronics instructor, I was forced to use a sub-standard textbook during
my first year of teaching. My students were daily frustrated with the many typographical errors
and obscure explanations in this book, having spent much time at home struggling to comprehend
the material within. Worse yet were the many incorrect answers in the back of the book to selected
problems. Adding insult to injury was the $100+ price.
Contacting the publisher proved to be an exercise in futility. Even though the particular text I
was using had been in print and in popular use for a couple of years, they claimed my complaint
was the first theyd ever heard. My request to review the draft for the next edition of their book
was met with disinterest on their part, and I resolved to find an alternative text.
Finding a suitable alternative was more difficult than I had imagined. Sure, there were plenty of
texts in print, but the really good books seemed a bit too heavy on the math and the less intimidating
books omitted a lot of information I felt was important. Some of the best books were out of print,
and those that were still being printed were quite expensive.
It was out of frustration that I compiled Lessons in Electric Circuits from notes and ideas I had
been collecting for years. My primary goal was to put readable, high-quality information into the
hands of my students, but a secondary goal was to make the book as affordable as possible. Over the
years, I had experienced the benefit of receiving free instruction and encouragement in my pursuit
of learning electronics from many people, including several teachers of mine in elementary and high
school. Their selfless assistance played a key role in my own studies, paving the way for a rewarding
career and fascinating hobby. If only I could extend the gift of their help by giving to other people
what they gave to me . . .
So, I decided to make the book freely available. More than that, I decided to make it open,
following the same development model used in the making of free software (most notably the various
UNIX utilities released by the Free Software Foundation, and the Linux operating system, whose
437

438

APPENDIX A-1.

ABOUT THIS BOOK

fame is growing even as I write). The goal was to copyright the text so as to protect my authorship
but expressly allow anyone to distribute and/or modify the text to suit their own needs with a
minimum of legal encumbrance. This willful and formal revoking of standard distribution limitations
under copyright is whimsically termed copyleft. Anyone can copyleft their creative work simply
by appending a notice to that effect on their work, but several Licenses already exist, covering the
fine legal points in great detail.
The first such License I applied to my work was the GPL General Public License of the
Free Software Foundation (GNU). The GPL, however, is intended to copyleft works of computer
software, and although its introductory language is broad enough to cover works of text, its wording
is not as clear as it could be for that application. When other, less specific copyleft Licenses began
appearing within the free software community, I chose one of them (the Design Science License, or
DSL) as the official notice for my project.
In copylefting this text, I guaranteed that no instructor would be limited by a text insufficient
for their needs, as I had been with error-ridden textbooks from major publishers. Im sure this book
in its initial form will not satisfy everyone, but anyone has the freedom to change it, leveraging my
efforts to suit variant and individual requirements. For the beginning student of electronics, learn
what you can from this book, editing it as you feel necessary if you come across a useful piece of
information. Then, if you pass it on to someone else, you will be giving them something better than
what you received. For the instructor or electronics professional, feel free to use this as a reference
manual, adding or editing to your hearts content. The only catch is this: if you plan to distribute
your modified version of this text, you must give credit where credit is due (to me, the original
author, and anyone else whose modifications are contained in your version), and you must ensure
that whoever you give the text to is aware of their freedom to similarly share and edit the text. The
next chapter covers this process in more detail.
It must be mentioned that although I strive to maintain technical accuracy in all of this books
content, the subject matter is broad and harbors many potential dangers. Electricity maims and
kills without provocation, and deserves the utmost respect. I strongly encourage experimentation
on the part of the reader, but only with circuits powered by small batteries where there is no risk of
electric shock, fire, explosion, etc. High-power electric circuits should be left to the care of trained
professionals! The Design Science License clearly states that neither I nor any contributors to this
book bear any liability for what is done with its contents.

A-1.2

The use of SPICE

One of the best ways to learn how things work is to follow the inductive approach: to observe
specific instances of things working and derive general conclusions from those observations. In
science education, labwork is the traditionally accepted venue for this type of learning, although
in many cases labs are designed by educators to reinforce principles previously learned through
lecture or textbook reading, rather than to allow the student to learn on their own through a truly
exploratory process.
Having taught myself most of the electronics that I know, I appreciate the sense of frustration
students may have in teaching themselves from books. Although electronic components are typically
inexpensive, not everyone has the means or opportunity to set up a laboratory in their own homes,
and when things go wrong theres no one to ask for help. Most textbooks seem to approach the task
of education from a deductive perspective: tell the student how things are supposed to work, then

A-1.3.

ACKNOWLEDGEMENTS

439

apply those principles to specific instances that the student may or may not be able to explore by
themselves. The inductive approach, as useful as it is, is hard to find in the pages of a book.
However, textbooks dont have to be this way. I discovered this when I started to learn a
computer program called SPICE. It is a text-based piece of software intended to model circuits and
provide analyses of voltage, current, frequency, etc. Although nothing is quite as good as building
real circuits to gain knowledge in electronics, computer simulation is an excellent alternative. In
learning how to use this powerful tool, I made a discovery: SPICE could be used within a textbook
to present circuit simulations to allow students to observe the phenomena for themselves. This
way, the readers could learn the concepts inductively (by interpreting SPICEs output) as well as
deductively (by interpreting my explanations). Furthermore, in seeing SPICE used over and over
again, they should be able to understand how to use it themselves, providing a perfectly safe means
of experimentation on their own computers with circuit simulations of their own design.
Another advantage to including computer analyses in a textbook is the empirical verification
it adds to the concepts presented. Without demonstrations, the reader is left to take the authors
statements on faith, trusting that what has been written is indeed accurate. The problem with
faith, of course, is that it is only as good as the authority in which it is placed and the accuracy
of interpretation through which it is understood. Authors, like all human beings, are liable to err
and/or communicate poorly. With demonstrations, however, the reader can immediately see for
themselves that what the author describes is indeed true. Demonstrations also serve to clarify the
meaning of the text with concrete examples.
SPICE is introduced early in volume I (DC) of this book series, and hopefully in a gentle enough
way that it doesnt create confusion. For those wishing to learn more, a chapter in the Reference
volume (volume V) contains an overview of SPICE with many example circuits. There may be more
flashy (graphic) circuit simulation programs in existence, but SPICE is free, a virtue complementing
the charitable philosophy of this book very nicely.

A-1.3

Acknowledgements

First, I wish to thank my wife, whose patience during those many and long evenings (and weekends!)
of typing has been extraordinary.
I also wish to thank those whose open-source software development efforts have made this endeavor all the more affordable and pleasurable. The following is a list of various free computer
software used to make this book, and the respective programmers:
GNU/Linux Operating System Linus Torvalds, Richard Stallman, and a host of others too
numerous to mention.
Vim text editor Bram Moolenaar and others.
Xcircuit drafting program Tim Edwards.
SPICE circuit simulation program too many contributors to mention.
TEX text processing system Donald Knuth and others.
Texinfo document formatting system Free Software Foundation.
LATEX document formatting system Leslie Lamport and others.

440

APPENDIX A-1.

ABOUT THIS BOOK

Gimp image manipulation program too many contributors to mention.


Appreciation is also extended to Robert L. Boylestad, whose first edition of Introductory Circuit
Analysis taught me more about electric circuits than any other book. Other important texts in
my electronics studies include the 1939 edition of The Radio Handbook, Bernard Grobs second
edition of Introduction to Electronics I, and Forrest Mims original Engineers Notebook.
Thanks to the staff of the Bellingham Antique Radio Museum, who were generous enough to let
me terrorize their establishment with my camera and flash unit.
I wish to specifically thank Jeffrey Elkner and all those at Yorktown High School for being willing
to host my book as part of their Open Book Project, and to make the first effort in contributing to its
form and content. Thanks also to David Sweet (website: (http://www.andamooka.org)) and Ben
Crowell (website: (http://www.lightandmatter.com)) for providing encouragement, constructive
criticism, and a wider audience for the online version of this book.
Thanks to Michael Stutz for drafting his Design Science License, and to Richard Stallman for
pioneering the concept of copyleft.
Last but certainly not least, many thanks to my parents and those teachers of mine who saw in
me a desire to learn about electricity, and who kindled that flame into a passion for discovery and
intellectual adventure. I honor you by helping others as you have helped me.
Tony Kuphaldt, July 2001
A candle loses nothing of its light when lighting another
Kahlil Gibran

Appendix A-2

CONTRIBUTOR LIST

A-2.1

How to contribute to this book

As a copylefted work, this book is open to revision and expansion by any interested parties. The
only catch is that credit must be given where credit is due. This is a copyrighted work: it is not
in the public domain!
If you wish to cite portions of this book in a work of your own, you must follow the same
guidelines as for any other copyrighted work. Here is a sample from the Design Science License:
The Work is copyright the Author. All rights to the Work are reserved
by the Author, except as specifically described below. This License
describes the terms and conditions under which the Author permits you
to copy, distribute and modify copies of the Work.
In addition, you may refer to the Work, talk about it, and (as
dictated by "fair use") quote from it, just as you would any
copyrighted material under copyright law.
Your right to operate, perform, read or otherwise interpret and/or
execute the Work is unrestricted; however, you do so at your own risk,
because the Work comes WITHOUT ANY WARRANTY -- see Section 7 ("NO
WARRANTY") below.
If you wish to modify this book in any way, you must document the nature of those modifications
in the Credits section along with your name, and ideally, information concerning how you may be
contacted. Again, the Design Science License:
Permission is granted to modify or sample from a copy of the Work,
producing a derivative work, and to distribute the derivative work
441

442

APPENDIX A-2.

CONTRIBUTOR LIST

under the terms described in the section for distribution above,


provided that the following terms are met:
(a) The new, derivative work is published under the terms of this
License.
(b) The derivative work is given a new name, so that its name or
title can not be confused with the Work, or with a version of
the Work, in any way.
(c) Appropriate authorship credit is given: for the differences
between the Work and the new derivative work, authorship is
attributed to you, while the material sampled or used from
the Work remains attributed to the original Author; appropriate
notice must be included with the new work indicating the nature
and the dates of any modifications of the Work made by you.
Given the complexities and security issues surrounding the maintenance of files comprising this
book, it is recommended that you submit any revisions or expansions to the original author (Tony R.
Kuphaldt). You are, of course, welcome to modify this book directly by editing your own personal
copy, but we would all stand to benefit from your contributions if your ideas were incorporated into
the online master copy where all the world can see it.

A-2.2

Credits

All entries arranged in alphabetical order of surname. Major contributions are listed by individual
name with some detail on the nature of the contribution(s), date, contact info, etc. Minor contributions (typo corrections, etc.) are listed by name only for reasons of brevity. Please understand
that when I classify a contribution as minor, it is in no way inferior to the effort or value of a
major contribution, just smaller in the sense of less text changed. Any and all contributions are
gratefully accepted. I am indebted to all those who have given freely of their own knowledge, time,
and resources to make this a better book!

A-2.2.1

Dennis Crunkilton

Date(s) of contribution(s): May 2005 to present


Nature of contribution: Nutmeg-spice plots added to chapters 4,5, & 6; 04/2006
Nature of contribution: Mini table of contents, all chapters except appedicies; html, latex,
ps, pdf; See Devel/tutorial.html; 01/2006.
Nature of contribution: Attenuator section to CH 1; completed CH 2 Theory; completed CH 3 Diodes added bibliography; 07/2007.
Contact at: dcrunkilton(at)att(dot)net

A-2.2.

CREDITS

A-2.2.2

443

Tony R. Kuphaldt

Date(s) of contribution(s): 1996 to present


Nature of contribution: Original author.
Contact at: liec0@lycos.com

A-2.2.3

Wayne Little

Date(s) of contribution(s): 2007


Nature of contribution: Author, Input to output phase shift, in Operational Amplifiers
chapter.
Contact at: http://forum.allaboutcircuits.com/member.php?u=8824

A-2.2.4

Warren Young

Date(s) of contribution(s): August 2002


Nature of contribution: Provided initial text for Power supply circuits section of chapter
9.

A-2.2.5

Your name here

Date(s) of contribution(s): Month and year of contribution


Nature of contribution: Insert text here, describing how you contributed to the book.
Contact at: my email@provider.net

A-2.2.6

Typo corrections and other minor contributions

line-allaboutcircuits.com (June 2005) Typographical error correction in Volumes 1,2,3,5,


various chapters ,(:s/visa-versa/vice versa/).
The students of Bellingham Technical Colleges Instrumentation program.
Clifford Bailey (July 2005) Correction, Ch 1.4, s/V/I equations in 13026.eps.
Colin Barnard (November 2003) Correction on the nationality of Alexander Graham Bell.
Michiel van Bolhuis (April 2007) Typo Ch 2, s/greater than tlectrons/greater that electrons..
Jeff DeFreitas (March 2006)Improve appearance: replace / and / Chapters: A1, A2.
Sean Donner (April 2005) Typographical error correction in Voltage and current section,
Chapter 4: BIPOLAR JUNCTION TRANSISTORS,(by a the/ by the) (no longer loaded/ no
longer unloaded).

444

APPENDIX A-2.

CONTRIBUTOR LIST

Sulev Eesmaa (February 2006) Correction, Ch 4, beta by definition- collector and emitter/
collector and base//.
Bill Heath (September-December 2002) Correction on illustration of atomic structure, and
corrections of several typographical errors.
Maciej Noszczyski (December 2003) Corrected spelling of Niels Bohrs name.
Dino Rosaldi (April 2006) Clarification, CH 4, concerning JFET Ohmic region.
Don Stalkowski (June 2002) Technical help with PostScript-to-PDF file format conversion.
Joseph Teichman (June 2002) Suggestion and technical help regarding use of PNG images
instead of JPEG.
Jered Wierzbicki (December 2002) Correction on diode equation: Boltzmanns constant
shown incorrectly.
Dave@allaboutcircuits.com (July 2007) Proofreading and correction of CH 1.
Derek@allaboutcircuits.com (July 2007) Correction, CH 2, s/reverse bias the base emitter
junction/reverse bias the base collector junction/.
Gogs@allaboutcircuits.com (July 2007) Correction of CH 2, s/usually error caused/error
caused measurement/.

Appendix A-3

DESIGN SCIENCE LICENSE


c 1999-2000 Michael Stutz stutz@dsl.org
Copyright
Verbatim copying of this document is permitted, in any medium.

A-3.1

0. Preamble

Copyright law gives certain exclusive rights to the author of a work, including the rights to copy,
modify and distribute the work (the reproductive, adaptative, and distribution rights).
The idea of copyleft is to willfully revoke the exclusivity of those rights under certain terms
and conditions, so that anyone can copy and distribute the work or properly attributed derivative
works, while all copies remain under the same terms and conditions as the original.
The intent of this license is to be a general copyleft that can be applied to any kind of work
that has protection under copyright. This license states those certain conditions under which a work
published under its terms may be copied, distributed, and modified.
Whereas design science is a strategy for the development of artifacts as a way to reform the
environment (not people) and subsequently improve the universal standard of living, this Design
Science License was written and deployed as a strategy for promoting the progress of science and
art through reform of the environment.

A-3.2

1. Definitions

License shall mean this Design Science License. The License applies to any work which contains
a notice placed by the works copyright holder stating that it is published under the terms of this
Design Science License.
Work shall mean such an aforementioned work. The License also applies to the output of
the Work, only if said output constitutes a derivative work of the licensed Work as defined by
copyright law.
Object Form shall mean an executable or performable form of the Work, being an embodiment
of the Work in some tangible medium.
445

446

APPENDIX A-3.

DESIGN SCIENCE LICENSE

Source Data shall mean the origin of the Object Form, being the entire, machine-readable,
preferred form of the Work for copying and for human modification (usually the language, encoding
or format in which composed or recorded by the Author); plus any accompanying files, scripts or
other data necessary for installation, configuration or compilation of the Work.
(Examples of Source Data include, but are not limited to, the following: if the Work is an
image file composed and edited in PNG format, then the original PNG source file is the Source
Data; if the Work is an MPEG 1.0 layer 3 digital audio recording made from a WAV format audio
file recording of an analog source, then the original WAV file is the Source Data; if the Work was
composed as an unformatted plaintext file, then that file is the the Source Data; if the Work was
composed in LaTeX, the LaTeX file(s) and any image files and/or custom macros necessary for
compilation constitute the Source Data.)
Author shall mean the copyright holder(s) of the Work.
The individual licensees are referred to as you.

A-3.3

2. Rights and copyright

The Work is copyright the Author. All rights to the Work are reserved by the Author, except
as specifically described below. This License describes the terms and conditions under which the
Author permits you to copy, distribute and modify copies of the Work.
In addition, you may refer to the Work, talk about it, and (as dictated by fair use) quote from
it, just as you would any copyrighted material under copyright law.
Your right to operate, perform, read or otherwise interpret and/or execute the Work is unrestricted; however, you do so at your own risk, because the Work comes WITHOUT ANY WARRANTY see Section 7 (NO WARRANTY) below.

A-3.4

3. Copying and distribution

Permission is granted to distribute, publish or otherwise present verbatim copies of the entire Source
Data of the Work, in any medium, provided that full copyright notice and disclaimer of warranty,
where applicable, is conspicuously published on all copies, and a copy of this License is distributed
along with the Work.
Permission is granted to distribute, publish or otherwise present copies of the Object Form of
the Work, in any medium, under the terms for distribution of Source Data above and also provided
that one of the following additional conditions are met:
(a) The Source Data is included in the same distribution, distributed under the terms of this
License; or
(b) A written offer is included with the distribution, valid for at least three years or for as long
as the distribution is in print (whichever is longer), with a publicly-accessible address (such as a
URL on the Internet) where, for a charge not greater than transportation and media costs, anyone
may receive a copy of the Source Data of the Work distributed according to the section above; or
(c) A third partys written offer for obtaining the Source Data at no cost, as described in paragraph (b) above, is included with the distribution. This option is valid only if you are a noncommercial party, and only if you received the Object Form of the Work along with such an offer.

A-3.5.

4. MODIFICATION

447

You may copy and distribute the Work either gratis or for a fee, and if desired, you may offer
warranty protection for the Work.
The aggregation of the Work with other works which are not based on the Work such as but
not limited to inclusion in a publication, broadcast, compilation, or other media does not bring
the other works in the scope of the License; nor does such aggregation void the terms of the License
for the Work.

A-3.5

4. Modification

Permission is granted to modify or sample from a copy of the Work, producing a derivative work,
and to distribute the derivative work under the terms described in the section for distribution above,
provided that the following terms are met:
(a) The new, derivative work is published under the terms of this License.
(b) The derivative work is given a new name, so that its name or title can not be confused with
the Work, or with a version of the Work, in any way.
(c) Appropriate authorship credit is given: for the differences between the Work and the new
derivative work, authorship is attributed to you, while the material sampled or used from the Work
remains attributed to the original Author; appropriate notice must be included with the new work
indicating the nature and the dates of any modifications of the Work made by you.

A-3.6

5. No restrictions

You may not impose any further restrictions on the Work or any of its derivative works beyond those
restrictions described in this License.

A-3.7

6. Acceptance

Copying, distributing or modifying the Work (including but not limited to sampling from the Work
in a new work) indicates acceptance of these terms. If you do not follow the terms of this License,
any rights granted to you by the License are null and void. The copying, distribution or modification
of the Work outside of the terms described in this License is expressly prohibited by law.
If for any reason, conditions are imposed on you that forbid you to fulfill the conditions of this
License, you may not copy, distribute or modify the Work at all.
If any part of this License is found to be in conflict with the law, that part shall be interpreted
in its broadest meaning consistent with the law, and no other parts of the License shall be affected.

A-3.8

7. No warranty

THE WORK IS PROVIDED AS IS, AND COMES WITH ABSOLUTELY NO WARRANTY,


EXPRESS OR IMPLIED, TO THE EXTENT PERMITTED BY APPLICABLE LAW, INCLUDING BUT NOT LIMITED TO THE IMPLIED WARRANTIES OF MERCHANTABILITY OR
FITNESS FOR A PARTICULAR PURPOSE.

448

APPENDIX A-3.

A-3.9

DESIGN SCIENCE LICENSE

8. Disclaimer of liability

IN NO EVENT SHALL THE AUTHOR OR CONTRIBUTORS BE LIABLE FOR ANY DIRECT,


INDIRECT, INCIDENTAL, SPECIAL, EXEMPLARY, OR CONSEQUENTIAL DAMAGES (INCLUDING, BUT NOT LIMITED TO, PROCUREMENT OF SUBSTITUTE GOODS OR SERVICES; LOSS OF USE, DATA, OR PROFITS; OR BUSINESS INTERRUPTION) HOWEVER
CAUSED AND ON ANY THEORY OF LIABILITY, WHETHER IN CONTRACT, STRICT LIABILITY, OR TORT (INCLUDING NEGLIGENCE OR OTHERWISE) ARISING IN ANY WAY
OUT OF THE USE OF THIS WORK, EVEN IF ADVISED OF THE POSSIBILITY OF SUCH
DAMAGE.
END OF TERMS AND CONDITIONS
[$Id:

dsl.txt,v 1.25 2000/03/14 13:14:14 m Exp m $]

Index
ratio, 207, 241
ratio, 180, 241
10-50 milliamp signal, 334
4-20 milliamp signal, 334
4-layer diode, 286
741 operational amplifier, 316

Beam power tube, 412


Bel, 7
Beta ratio, 180, 241
Beta ratio, bipolar transistor, 417
Beta variations, 181
Bias current, op-amp, 353
Bias, diode, 91
Bias, transistor, 191, 213
Bilateral, 260
Bipolar-mode MOSFET, 279
Bistable, 344
Brattain, Walter, 56, 61
Breakdown, diode, 95
Breakdown, transistor, 291
Breakover, thyristor, 291
Bridge rectifier circuit, 104
Bridge rectifier circuit, polyphase, 105
Bypass capacitor, 238

A-weighted dB scale, 14
A/D converter, 321
AC-DC power supply schematic, 296
Active device, 2
Active mode, transistor, 176
Alpha ratio, 207, 241
Amplification, definition, 3
Amplifier, differential, 313
Amplifier, inverting, 327
Amplifier, noninverting, 326
Amplifier, single-ended, 313
Analog-to-digital converter, 321
Angular Momentum quantum number, 30
Anti-static foam, 253
Antilogarithm, 9
Artifact, measurement, 391
Astable, 346
Attenuator, 15
Attenuator, bridged T, 20
Attenuator, coaxial, 22
Attenuator, L, 20
Attenuator, PI, 19
Attenuator, rf, 22
Attenuator, T, 18
Avalanche photodiode, 145
Averager, 335

Calculus, 316, 341, 379


Capacitance, diode, 101
Capacitor, bypass, 238
Capacitor, coupling, 221
Capacitor, op-amp compensation, 359
Cat-Eye tube, 424
Cathode, 408
Cathode Ray Tube, 423
Center-tap rectifier circuit, 103
Characteristic curves, transistor, 180, 259
Check valve, 92
Clamper circuit, 114
Class A amplifier operation, 213
Class AB amplifier operation, 215
Class B amplifier operation, 214
Class C amplifier operation, 216
Class D amplifier operation, 216
Class, amplifier operation, 213

Band, electron, 44
Bandwidth, amplifier, 233
Bardeen, John, 56, 61
449

450
Clipper circuit, 110
clipper, zener diode, 134
CMRR, 348
Cockcroft-Walton, voltage multiplier, 121
Coherent light, 143
Cold-cathode tube, 420
COMFET, 279
Common-base amplifier, 207
Common-collector amplifier, 199
Common-emitter amplifier, 185
Common-mode rejection ratio, 348
Common-mode voltage, 348
Commutating diode, 123, 124
Commutation, 124
Commutation time, diode, 101
Commutation, forced, 307, 308
Commutation, natural, 297, 308
Comparator, 318
Compensation capacitor, op-amp, 359
Conduction band, 44
Conductivity-Modulated Field-Effect Transistor, 279
Constant-current diode, 154, 183
Controlled rectifier, 301
Conventional flow, 91
Cooper pair, 76
Coupling capacitor, 221
Coupling loop, resonator, 426, 429
Critical rate of voltage rise, 291, 293
Crossover distortion, 432
Crowbar, 296
CRT, 423
Crystal radio, 371
Current mirror, 241
Current source, 177, 333
Current sourcing vs. sinking, 243
Current, diode leakage, 101
Current-limiting diode, 154
Current-regulating diode, 154
Curve, characteristic, 180, 259
Cutoff voltage, 251
Cutoff, transistor, 169, 176
Czochralski process, silicon , 71
Darlington pair, 205
Datasheet, component, 100

INDEX
dB, 8
dB, absolute power measurements, 14
dB, sound measurements, 14
dBA, 14
dBk, 14
dBm, 14
dBW, 14
DC restorer circuit, 114
Decibel, 8
Decibels, attenuator, 16
Decineper, 12
Degenerative feedback, 233
Derivative, calculus, 380
DIAC, 292
Differential amplifier, 313
Differential pair, 363, 364
Differentiation, 316
Differentiation, calculus, 341, 379
Diode, 91
Diode check, meter function, 97, 172
Diode equation, the, 95
Diode junction capacitance, 101
Diode leakage current, 101
Diode PIV rating, 95
Diode tube, 408
Diode, 4-layer, 69
Diode, constant-current, 154, 183
Diode, Esaki, 137
Diode, four-layer, 286
Diode, hot carrier, 135, 151
Diode, IMPATT, 152
Diode, laser, 143
Diode, light-activated, 144
Diode, light-emitting, 138
diode, MIIM, 80
Diode, pin, 152
Diode, PNPN, 286
Diode, schottky, 135
Diode, Shockley, 286
Diode, snap, 151
Diode, SPICE, 155
Diode, tunnel, 136
Diode, varactor, 151
Diode, varicap, 151
Diode, zener, 128
DIP, 317

INDEX

451

Discharge tube, 421


Distortion, amplifier, 233
Distortion, crossover, 432
dn, 12
Double-layer tunneling transistor, 80
Drift, op-amp, 358
Dropout, thyristor, 291
Dual Inline Package, 317
Dual power supply, 312
Duty cycle, square wave, 319
Duty cycle, squarewave, 217

Gain, AC versus DC, 6


Gate turn off switch, 69
Gate-Controlled Switch, 293
Gate-Turn-Off thyristor, 293
GCS, 293
Glow tube, 422
Grid, 407
Ground, 312
Ground, virtual, 327
GTO, 293
GTO, gate turn off switch, 69

Edison effect, 407


Effect, Edison, 407
Electrode, cathode, 408
Electrode, grid, 407
Electrode, screen, 411
Electrode, suppressor, 413
Electron, 26
Electron flow, 91
Emitter follower, 202
Equation, diode, 95
Equilibrium, 322
Esaki diode, 137
Exclusion principle, 33

Half-wave rectifier circuit, 102


Harmonic, 304
Harmonic, even vs. odd, 304
Harmonics and waveform symmetry, 304
Heptode, 414
hfe, 181
High temperature superconductors:, 78
Holding current, SCR, 295
hot carrier diode, 135
Hot-cathode tube, 420
Hybrid parameters, 181
Hysteresis, 344, 420

Failure mode, zener diode, 129


Faradays Law, 122, 124
Feedback, amplifier, 233
Feedback, negative, 322
Feedback, positive, 343
FET, field effect transistor, 61
Field effect transistor , 61
Firing, thyristor, 291
Flash converter, 321
Floating, 169, 293
Flow, electron vs. conventional, 91
Foam, anti-static, 253
Forced commutation, 307, 308
Forward bias, 91
Forward voltage, diode, 93
Four-layer diode, 286
Frequency response, op-amp, 359
Full-wave rectifier circuit, 103, 104
Gain, 6

IC, 243
IGBT, 279, 310
IGFET, insulated gate field effect transistor,
66
IGT, 279, 310
IMPATT diode, 152
Inductive output tube, 426
Inert elements, 36
Input, inverting, 313
Input, noninverting, 313
Insulated gate field effect transistor, 66
Insulated-Gate Bipolar Transistor, 279, 310
Insulated-Gate Transistor, 279, 310
Integrated circuit, 243
Integration, calculus, 341, 379
Inverting amplifier, 187, 327
Inverting summer, 336
Ionization, 282, 419
JFET, junction field effect transistor, 61
Josephson junctions, 76

452
Josephson transistor, 76
Joules Law, 11, 129
Junction capacitance, diode, 101
Kickback, inductive, 122
Kirchhoffs Current Law, 167
Kirchhoffs Voltage Law, 202
Klystron, 426
Laser diode, 143
Laser light, 143
Latch-up, 351
Latching, thyristor, 291
Leakage current, diode, 95, 101
LED, 138
Light-emitting diode, 138
Lilienfeld, Julius, 61
Load line, 217
Logarithm, 9
Magic-Eye tube, 424
Magnetic quantum number, 30
Magnetic tunnel junction, 84
Mechanics, quantum, 29
MESFET, metal semiconductor field effect transistor, 65
Metal oxide field effect transistor, 66
Mho, 262
Microphonics, electron tube, 431
MIIM, diode, 80
Monochromatic light, 143
MOS Controlled Thyristor, 309
MOS-gated thyristor, 309
MOSFET, metal oxide field effect transistor,
66
MTJ, magnetic tunnel junction, 84
Mu, tube amplification factor, 415
Multiplier circuit, diode, 116
Natural commutation, 297, 308
Negative feedback, 233, 322
Negative resistance, 136
Neper, 12
Neutron, 26
Noble elements, 36
Noninverting amplifier, 326

INDEX
Noninverting summer, 336
Number, quantum, 30
Offset null, op-amp, 352
Offset voltage, op-amp, 351
Ohmic region, JFET, 261
Op-amp, 239, 316
Operational amplifier, 239, 316
Orbital, electron, 32
Oscillator, 233
Oscillator, op-amp, 346
Oscillator, relaxation, 283
Oscillator, voltage-controlled, 427
Over-unity machine, 3
Passive averager, 335
Passive device, 2
Pauli, exclusion principle, 33
PCB, 99
Peak detector, 108
Pentagrid tube, 414
Pentode tube, 271
Perpetual motion machine, 3
Phase shift, op-amp, 360
Photodiode, 144
Photodiode amplifier, 396
Photodiode, APD, 145
Photodiode, PIN, 145
PI-network , 16
PIN diode, 152
PIN, photodiode, 145
Pinch-off voltage, 251
PIV rating, diode, 95
Plasma, 282, 419
PNPN diode, 286
Polyphase bridge rectifier circuit, 105
Positive feedback, 233, 281, 343
Power supply schematic, AC-DC, 296
Principal quantum number, 30
Printed circuit board, 99
Process variable, 315
Proton, 26
Pulse-width modulation, 320
Push-pull amplifier, 214
PWM, 320
Quantum dot, 82

INDEX
Quantum dot transistor, 82
Quantum mechanics, 29
Quantum number, 30
Quantum physics, 25
quantum tunneling, 79
Quiescent, 217
Radio, crystal, 371
Rail voltage, 324
Rectifier, 91
Rectifier circuit, 102
Rectifier circuit, full-wave, 103, 104
Rectifier circuit, half-wave, 102
Rectifier, controlled, 301
Reference junction, thermocouple, 353
Reflex klystron, 427
Regenerative feedback, 233, 281
Regulator, voltage, 204
Relaxation oscillator, 283
Resistance, negative, 136
Resonant tunneling diode, 79
Restorer circuit, 114
Reverse bias, 91
Reverse recovery time, diode, 101
Reverse voltage rating, diode, 95
Rheostat, 182, 262
Richter scale, 8
Ripple voltage, 106
Runaway, thermal, 236
s,p,d,f subshell notation, 32
Saturable reactor, 2
Saturation voltage, 324
Saturation, transistor, 169, 176
Schottky diode, 135
SCR, 293, 423
SCR bridge rectifier, 301
SCR, silicon controlled rectifier, 69
Screen, 411
SCS, 307
Secondary emission, 412
Semiconductor, defined, 45
Sensitive gate, SCR, 295
Setpoint, 315
Shell, electron, 31
Shockley diode, 286

453
Shockley, William, 56, 61, 69
Siemens, 262, 334
Signal, 10-50 milliamp, 334
Signal, 4-20 milliamp, 334
Silicon controlled rectifier, 69
Silicon-controlled rectifier, 293, 423
Silicon-controlled switch, 307
Single-ended amplifier, 313
Sink, current, 243
Slicer circuit, 110
Slide rule, 9
Small-scale integration, 365
Snap diode, 151
Snubber, 124
Solar cell, 146
Solid-state, 2
Sound intensity measurement, 14
Spark gap, 420
SPICE, diode, 155
Spin quantum number, 30
Spintronics, 83
Split power supply, 312
SQUID:, 76
SSI, 365
Step recovery diode, 151
Subshell notation, 32
Subshell, electron, 31
Superconduction quantum interference device,
76
Superconductivity, 75
Superposition theorem, 222
Suppressor, 413
Switching time, diode, 101
T-network , 16
Tetrode tube, 271, 411
Theorem, Superposition, 222
Thermal runaway, BJT, 236
Thermal voltage, diode, 95
Thermocouple, 353
Three-phase bridge rectifier circuit, 105
Thyratron, 423
Thyratron tube, 283
Thyristor, 69, 420
Time, diode switching, 101
Totalizer, 342

454
Transconductance, 262, 334
Transconductance amplifier, 334
Transistor, field effect, 61
Transistor, insulated gate field effect, 66
Transistor, Josephson, 76
Transistor, metal oxide field effect, 66
Transistor, single electron, 82
Triode tube, 271, 283, 408
Tube, discharge, 421
Tunnel diode, 136
Tunnel junction, magnetic, 84
tunneling, quantum, 79
Unipolar, conduction, 61
Unit, bel, 7
Unit, decineper, 12
Unit, mho, 262
Unit, neper, 12
Unit, siemens, 262, 334
Valence band, 44
Valence shell, 32
Valve, check, 92
Varactor diode, 151
Varicap diode, 151
VCO, 285
Virtual ground, 327
VMOS transistor, 279
Voltage buffer, 324
Voltage doubler circuit, 116
Voltage follower, 202, 324
Voltage multiplier circuit, 116
Voltage multiplier, Cockcroft-Walton, 121
Voltage regulator, 204
Voltage regulator tube, 422
Voltage rise, critical rate of, 291, 293
Voltage, bias, 191, 213
Voltage, common-mode, 348
Voltage, forward, 93
Voltage, op-amp output saturation, 324
Voltage, ripple, 106
Voltage-controlled oscillator, 285, 427
Volume units, 14
VU scale, 14
Waveform symmetry and harmonics, 304

INDEX
Zener diode, 128
Zener diode failure mode, 129
Zener diode, clipper, 134

INDEX
.

455

Anda mungkin juga menyukai