28.01.2015 Views

1 - Andrzej Materka

1 - Andrzej Materka

1 - Andrzej Materka

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

POLITECHNIKA L6DZKA<br />

WYDZIAt ELEKTROTECHNIKI I ELEKTRONIKI<br />

<strong>Andrzej</strong> <strong>Materka</strong><br />

Analog Electronics<br />

Lecture notes<br />

L6dz,1998


QM- 0041 ml&oif<br />

Recenzent / Reviewer<br />

Prof, dr in. Jersgi Ludnski<br />

Sklad komputerowy, przygotowanie rysunkow i przelamanie tekstu /<br />

Computer text-editing, figure artworks, text formatting<br />

Andrgj <strong>Materka</strong><br />

Projekt graficzny okladki / Cover graphics design<br />

Wojtek <strong>Materka</strong><br />

© Copyright by Wydzial Elektrotechniki i Elektroniki Politechniki Lodzkiej<br />

ISBN 83-87202-06-1<br />

Na okladce: Seria dziewi^ciu analogowych ukladow scalonych, z ktorych kazdy pelni<br />

funkcje. sztucznej sieci neuronowej Kohonena. Uklady zaprojektowano w zespole<br />

naukowo-badawczym kierowanym przez Autora w Instytucie Elektroniki Politechniki<br />

Lodzkiej, we wspolpracy z prof. Alexisem De Vosem z Uniwersytetu w Gandawie.<br />

Zostaly one wyprodukowane przez fabryke. ukladow scalonych 'Alcatel' w Belgii. Uklad<br />

sieci Kohonena pozwala na praktyczna. realizacje. idei uczenia nienadzorowanego, m. in.<br />

do celow rozpoznawania obrazow.<br />

Front cover: A lot of 9 analog integrated circuits, each performing the function of<br />

Kohonen-type artificial neural network. The research team led by the Author in the<br />

Institute of Electronics, Technical University of Lodz - in cooperation with Professor<br />

Alexis De Vos of Gent University - has designed the circuits. The 'Alcatel' silicon<br />

foundry in Belgium fabricated the ICs. The Kohonen network implements the idea of<br />

unsupervised learning for pattern recognition - among many other applications.


Contents<br />

1 Introduction, 9<br />

1.1 Signals and Their Spectra, 9<br />

1.2 Analog and Digital Signals, 12<br />

1.3 Electronic-System Block Diagrams, 15<br />

1.4 Information-Processing Electronics Versus Power Electronics, 16<br />

1.5 Behavior of Circuit Components Versus Frequency, 16<br />

1.6 Summary, 17<br />

2 Amplifiers: Behavioral Description, 18<br />

2.1 Basic Concepts, 18<br />

2.2 Cascaded Amplifiers, 22<br />

2.3 Power Supplies and Efficiency, 23<br />

2.4 Decibel Notation, 25<br />

2.5 Frequency Response, 26<br />

2.6 The Miller Theorem, 34<br />

2.7 Linear Distortion, 37<br />

2.8 Pulse Response, 39<br />

2.9 Nonlinear Distortion, 44<br />

2.10 Summary, 46<br />

3 Diode Circuits, 48<br />

3.1 The Ideal Diode, 48<br />

3.2 Terminal Characteristics of Semiconductor Diodes, 52<br />

3.3 Analysis of Diode Circuits, 56<br />

3.4 The Diode Small-Signal Model at Low Frequencies, 61<br />

3.5 Rectifier Circuits, 66<br />

3.6 Zener-Diode Voltage Regulator Circuits, 70<br />

3.7 Wave-Shaping Circuits, 73<br />

3.b Switching and High-Frequency Behavior of thepn Junction, 76<br />

3.9 Special Diodes, 84<br />

3.10 Summary, 86<br />

4 Field-Effect Transistor Circuits, 88<br />

4.1 The »-channel Junction FET, 88<br />

4.2 Metal-Oxide-Semiconductor FETs, 91<br />

4.3 Load-Line Analysis of a Simple JFET Amplifier, 94<br />

4.4 The Self-Bias Circuit, 96<br />

4.5 The Fixed- Plus Self-Bias Circuit, 98<br />

4.6 The Small-Signal Equivalent Circuit, 100<br />

4.7 Basic Small-Signal FET Amplifier Circuits, 103<br />

4.8 The FET as a Voltage-Controlled Resistance, 113<br />

4.9 CMOS Analog Switch, 115<br />

4.10 CMOS Logic Circuits, 116<br />

4.11 FET Dynamic Circuit Model, 121<br />

4.12 Summary, 125<br />

-3-


5 Bipolar Transistor Circuits, 127<br />

5.1 Load-Line Analysis of a Common-Emitter Amplifier, 130<br />

5.2 The pnp Bipolar Junction Transistor, 133<br />

5.3 Secondary Effects, 135<br />

5.4 Large-Signal dc BJT Models, 139<br />

5.5 Large-Signal dc Analysis of BJT Circuits, 141<br />

5.6 Four-Resistor Bias Circuit, 145<br />

5.7 Small-Signal Equivalent Circuits, 149<br />

5.8 The Common-Emitter Amplifier, 150<br />

5.9 The Emitter Follower, 153<br />

5.10 Review of Small-Signal Equivalent Circuit Analysis, 156<br />

5.11 The Common-Emitter Hybrid-Parameter Small-Signal Model, 160<br />

5.12 The Hybrid-* Model, 161<br />

5.13 Bipolar Transistor Behavior at High Frequencies, 164<br />

5.14 Large-Signal Dynamic Model for die BJT, 173<br />

5.15 Switching Behavior of die BJT, 177<br />

5.16 Summary, 181<br />

6 Feedback Circuits, 183<br />

6.1 Effects on Sensitivity, Bandwidth and Distortion, 184<br />

6.2 Feedback Types, 194<br />

6.3 Effect of Feedback Types on Input and Output Impedance, 196<br />

6.4 Summary of the Effect of Various Feedback Types, 198<br />

6.5 Practical Feedback Networks, 199<br />

6.6 Stability of Feedback Amplifiers, 202<br />

6.7 Sinusoidal Oscillators, 205<br />

6.8 Summary, 221<br />

References, 223<br />

Review Questions, 224<br />

Problems, 230<br />

Appendix<br />

A. Nominal Values and the Color Code for Resistors, 256<br />

B. Introduction to PSpice, 258<br />

C. English-Polish Dictionary of Selected Terms, 266<br />

D. Manufacturers' Date Sheets for 2N2222A transistor, 272<br />

-4-


Przedmowa / Preface<br />

Niniejszy skrypt powstal w wyniku zebiania i uporzajikowania notatek z wykladow<br />

Analog Electronics oraz Electronic Devices and Systems, ktore mialem przyjemnosc<br />

prowadzic w Politechnice Lodzkiej dla kilku grup studenckich IFE (International Facility<br />

of Engineering) w latach 1995-1997. Opisuje. w nim zasady dzialania i wlasciwosci<br />

podstawowych ukiadow elektronicznych - analogowych.<br />

Zakres materialu opisanego w skrypcie w niewielkim jedynie stopniu obejmuje olbrzymia,<br />

rozmaitosc znanych ukiadow elektronicznych. Staralem si zawrzec w nim informacje<br />

podstawowe, zdefiniowac terminologie. i omowic elementarne obwody zawieraja.ce diody,<br />

tranzystory polowe, bipolarne a takze obwody ze sprzzeniem zwrotnym. Oprocz<br />

wiadomosci teoretycznych i przykladow symulacji komputerowych skrypt zawiera<br />

cwiczenia rachunkowe, ktore pozwalajq. na ilosciowe zilustrowanie wlasciwosci ukiadow,<br />

a w niektorych przypadkach lepsze zrozumieiiie zalozeri projektowych. Pytania<br />

powtorkowe zwracaj% uwage na zasadnicze elementy tresci merytorycznej<br />

poszczegolnych rozdzialow i moga. bye pomocne w porza_dkowaniu wiedzy przed<br />

egzaminem.<br />

Program studiow inzynierskich w IFE zawiera zmniejszona^ liczbe. godzin zajee z<br />

ukiadow elektronicznych w porownaniu ze studiami magisterskimi na Wydziale<br />

Elektrotechniki i Elektroniki PL. Dla lepszego wykorzystania przydzielonych limitow<br />

czasowych i zwikszenia skutecznosci przekazu wiedzy, wyklady swoje ilustrowaiem<br />

dose obficie przyktadami obhezen za pomoca. programu SPICE. W tym celu<br />

prowadzilem wyklady w laboratorium komputerowym, w ktorym kazdy ze studentow<br />

mogl samodzielnie realizowac symulacje omawianych ukiadow. Ta forma zajec<br />

dydaktycznych dowiodla swojej uzytecznosci i jest godna polecenia innym,<br />

zainteresowanym wykladowcom. Przyklady wybranych programow w jzyku SPICE<br />

zawarlem w skrypcie. Zachecam Czytelnikow do korzystania z tych przykladow, a takze<br />

do pisania wlasnych programow i wykonywania symulacji ukiadow. Symulacje takie nie<br />

moga. zastapic, bez wa_tpienia, doswiadczeri praktycznych i pomiarow rzeczywistego<br />

obwodu. Sa_ one jednak mniej kosztowne, trwaj% krocej, a w przypadku projektowania<br />

ukiadow scalonych sa. jedyna. forma, sprawdzania poprawnosci projektu przed faza.<br />

realizacji. Zainteresowanym mog udostpnic darmowa. edukacyjna_ wersje. programu<br />

PSpice (dwie dyskietM 3.25"). Przy tej okazji chcialbym podziekowac Wydawnictwu<br />

Prentice Hall, ktore przyslalo mi ten program wraz z ksia^zka. [12] do oceny, kiedy<br />

pracowalem jako wykladowca w Monash University w Australii.<br />

Z faktu, iz studenci IFE — sluchacze moich wykladow — z niecierpliwoscia. pytali o<br />

kolejne porcje notatek wnosz, ze napisanie skryptu bylo potrzebne. Wypelnia on<br />

okreslona. iuk w krajowym pismiennictwie akademickim, dotyczaca. zwlaszcza<br />

materialow optacowanych w jezykach obcych, co jest szczegolnie wazne w okresie<br />

otwierania si Polski na swiat Z drugiej strony, polskojzyczni studenci IFE powinm<br />

miec mozliwosc zapoznania si z rodzima. terminologie w dziedzinie ich studiow. Na ich<br />

potrzeby opracowalem slownik angielsko-polski terminow uzywanych w skrypcie,<br />

zawarty w Dodatku. Trzeba podkreslic, ze lektura niniejszego skryptu nie moze zastaj>ic<br />

prawdziwie poglbionych studiow przedmiotu. W tym celu odsylam Czytelnikow do<br />

-5-


dziel obszetniejszych, z ktotych kotzystalem przygotowujajc notatki do moich wykladow.<br />

List tych prac zawieta spis literatury.<br />

Na koniec chc zaznaczyc, ze optacowanie niniejsze nie mogloby powstac gdyby moja<br />

Zona i moj Syn nie zgodzili si na to, abym 30 weekendow toku 1997 spdzil z edytotem<br />

tekstu. Setdecznie Im za to dziekuj. Jestem towniez wdziczny Dziekanowi Wydzialu<br />

Elekttotechniki i Elekttoniki Politechniki Lodzkiej, Panu Ptofesotowi Doktotowi Janowi<br />

Leszczyriskiemu za ufundowanie honotatium, ktote pozwoli mi na modetnizacj mojego<br />

komputeta. Dziki temu zmniejsz troch dystans do czolowki tego fascynuja.cego i<br />

ptzeiazaja_cego wyscigu powszechniej komputetyzacji, u ktoiej podstaw znajduje si<br />

postp w dziedzinie elekttoniki. No doubts it is...<br />

Zycz Panstwu ptzyjemnej i owocnej lektuty -<br />

<strong>Andrzej</strong> <strong>Materka</strong>.<br />

Lodz, wlutym 1998.


Preface / Przedmowa<br />

This textbook is a systematized collection of the notes of lectures that I had a pleasure to<br />

deliver in 1995-1997 to a number of student groups at the International Faculty of<br />

Engineering (IFE), Technical University of Lodz. It describes the principle of operation<br />

and properties of basic analog electronic circuits.<br />

The material included in this textbook covers the huge variety of known electronic<br />

circuits to a very limited degree only. I have made efforts to introduce fundamental<br />

concepts, define terminology, and discuss elementary circuits that comprise diodes, fieldeffect<br />

and bipolar transistors, as well as circuits with feedback. Besides the theory and<br />

computer simulation examples, the lecture notes include problems and exercises for the<br />

student, which help quantitatively illustrate various circuit properties and provide better<br />

understanding of design specifications in some cases. Review questions'provided turn the<br />

reader's attention to essential elements of the theory presented in individual chapters;<br />

they can be helpful to sort things out before exams.<br />

The 4-year BEng study program at IFE allows less class time for Analog Electronics than<br />

the 5-year MSc studies at the TUL's Faculty of Electrical and Electronic Engineering do.<br />

To make better use of the student contact-hours allocated, I kept amply illustrating my<br />

lectures with computer simulation examples, using the SPICE program. To achieve that,<br />

a computer laboratory was chosen to be the place to give the lectures, where almost<br />

every listener had an access to a PC terminal to individually exercise the virtual<br />

experiments with the circuits under study. This form of carrying on the teaching has<br />

proven its usefulness and deserves recommendation to other interested lecturers. I<br />

integrated examples of the SPICE code with the text of the lecture notes. I encourage<br />

the readers to use these examples and to write their own programs for circuit simulation.<br />

Simulations can not, of course, replace practical experiments and measurements of<br />

physical circuits. They are, however, less expensive, last shorter, and in the case of<br />

integrated circuits are the only means of verifying the circuit correctness before its<br />

fabrication. The interested readers can copy a free educational version of the PSpice<br />

program (two 3.5 inches floppies) from my resources. On this occasion I wish to thank<br />

the Prentice Hall publishing house representative who sent me this program along with<br />

the book [12] for evaluation when I was with Monash University, Melbourne, Australia,<br />

working as a lecturer.<br />

From the fact that IFE students - the listeners to my lectures — kept impatiently asking<br />

me about consecutive printouts of the lecture notes I gather writing the textbook was<br />

needed. It fills up a gap in the country's academic literature that concerns especially<br />

foreign-language material that is particularly important in the period of Poland's opening<br />

to the outside world. On the other hand, those of students who are Polish native<br />

speakers should be given an opportunity to familiarize with their mother-tongue<br />

terminology in the field of their studies. I then elaborated a short English-Polish<br />

dictionary of selected terms utilized in the textbook, which is included in the Appendix.<br />

It should be emphasized that reading this textbook can not replace truly thorough studies<br />

of ^he subject. To do such studies, I recommend the readers to refer to more extensive<br />

-7-


ooks that I used to prepare the lecture notes. The list of these works is provided at the<br />

end of the text<br />

Finally, I wish to point out that this textbook would not appear if my wife and my son<br />

would not agree that I might spend 30 week-ends in 1997 with the text editor. I do<br />

sincerely thank them for this opportunity. I am also grateful to the Dean of the Faculty<br />

of Electrical and Electronic Engineering, Professor Jan Leszczynski, for founding a<br />

scholarship that will help upgrade my computer. Then I will be able to cut a litde bit die<br />

distance to the leaders of the fascinating and frightening race of widespread<br />

computerization that is actually based on the progress in electronics. Bez wa.tpienia tak<br />

jest...<br />

I wish you enjoyable and fruitful reading,<br />

Lodz, in February 1998.<br />

<strong>Andrzej</strong> <strong>Materka</strong>.


Analog Electronics /Introduction<br />

1. Introduction<br />

In these lecture notes we shall study electronic devices and their interconnections that form discrete<br />

or integrated circuits (ICs). These circuits are devised to perform a variety of functions within<br />

electronic systems' of interest The well-known examples of contemporary electronic systems are<br />

Walkman radio, TV set, PC computer, TV satellite receiver, computerized monitors for patients in<br />

intensive-care units, cellular phones, and the like. There exist also a variety of complex systems,<br />

which do not operate based mainly on electrons' movement but nevertheless would not function<br />

properly without an embedded electronic circuitry. These include aircraft, photocopying machines,<br />

weather satellites, etc. Electronic systems control fuel mixture and ignition timing to maximize<br />

performance and minimize undesirable emissions from automobile engines. It is in fact impossible<br />

to overestimate the role of electronics in the modem society. Perhaps there is no human activity<br />

area, ranging from daily life to space exploration, in which we don't utilize the electronic devices.<br />

The foundations of electronics were established by the observations of M. Faraday (1791-1867) and<br />

the discovery of electron by J. J. Thomson in 1897, but its development proceeded relatively slowly<br />

until the importance of radar became apparent at the beginning of World War II. During that<br />

period and extending until about 1955, the expanding field of electronics depended heavily on the<br />

principle of electron emission from the hot cathode of a vacuum tube. By modern standards,<br />

vacuum tube electronics was expensive, bulky, hot, unreliable and even dangerous because of the<br />

high voltages present in the tubes. With the discovery of the transistor by J. Bardeen, W. H. Brittain<br />

and W. B. Schockley in 1948, the stage was set for the solid-state electronics explosion of today.<br />

This technological explosion is often compared to the industrial revolution.<br />

Since the invention of transistor the technology has progressed from an individually prepared<br />

laboratory device (a tube) to millions of components (diodes, transistors) on a single silicon wafer<br />

only a millimeter a side. Coupled with this astounding miniaturization has been a corresponding<br />

decrease in the price per component For example, a current popular microcomputer chip,<br />

containing more than one million components, costs about the same as one of the electron tubes of<br />

earlier electronics. At the same time, with the increased internal complexity and overall functionality<br />

of solid- 1 tate chips, the number of external components is highly reduced as compared to vacuum<br />

tube electronic circuits. Also, the temperature of device operation is decreased. These two main<br />

factors lead to a much higher reliability of solid state electronic systems.<br />

1.1 Signals and Their Spectra<br />

The main function of most electronic circuits is either to process or to generate signals. Signals<br />

contain information about a variety of things in our physical world. For example, information about<br />

the weather is contained in signals that represent the air temperature, pressure, the wind speed, etc.<br />

The voice of a radio announcer reading the news into the microphone provides an acoustic signal<br />

that carries information about world affairs. To monitor and diagnose the status of the human<br />

body, biopotentials are measured on the surface of the skin to represent the activities of the heart<br />

(ECG - the electrocardiogram), the brain (EEG - electroencephalogram), the muscles (EMG -<br />

electromyogram) and others.<br />

Processing of signals is needed to extract information from them, for an observer (be it a human or<br />

a machine). Electronic circuits and systems most conveniently perform the signal processing. For<br />

this to be possible, the signals, which are not originally in the form of an electrical signal, must first<br />

-9-


Analog Electronics /Introduction<br />

be converted into a voltage or current Devices known as transducers accomplish this process. A<br />

variety of transducers exist, each suitable for one of the various forms of physical signals. For<br />

instance, a microphone is in effect a pressure transducer. It is not our purpose here to study<br />

transducers; rather, we shall assume that the signals of interest already exist in the electrical domain.<br />

For the purpose of our study the signals will be represented by one of the equivalent forms shown<br />

in Fig. 1.1. In Fig. 1.1(a) the signal is represented by a voltage source v//) having a source resistance<br />

R r In the alternate representation of Fig. 1.1(b), the signal is represented by a current source /,(/)<br />

having a source resistance R r Although the two representations are equivalent, the one shown in<br />

Fig. 1.1(a) (known as the Thevenin form) is preferred when R, is low. The representation of Fig.<br />

1.1(b) (known as the Norton form) is preferred when R s is high.<br />

From the discussion above it should be apparent that a signal is a time-varying quantity that can be<br />

represented by a graph such as those shown in Fig. 1.2. In fact, the information content is<br />

represented in changes in signal magnitude as time progresses. The overall level of the magnitude is<br />

one of the signal characteristic features. Table 1.1 gives examples of the range of magnitudes of<br />

different signals. In general, waveforms are difficult to characterize mathematically. In other words,<br />

it is difficult to describe succincdy an arbitrary looking waveform, such as one of those of Fig. 1.2.<br />

TABLE 1.1<br />

MAGNITUDE AND INTERNAL RESISTANCE OF SELECTED SIGNAL SOURCES<br />

SIGNAL SOURCE<br />

RANGE OF INTERNAL RESISTANCE<br />

MAGNITUDES<br />

ECG


Analog Electronics /Introduction<br />

The essential parts of the spectra of practical signals are usually confined to relatively shon<br />

segments of the frequency axis. For instance, die spectrum of audible sounds such as speech and<br />

music extends from about 20Hz to about 20kHz - a frequency range known as die audio band.<br />

This observation is very useful in the processing of such signals and important for electronic circuit<br />

designers. This is because the properties of signal processing circuits have to match the signal<br />

properties. For example, there is no point in amplifying the signal in the frequency range in which<br />

there is no signal component of significant energy. This would mean amplification of noise and<br />

interference, which corrupt the information. Sometimes the range of frequencies occupied by a<br />

signal can be changed. In radio communication, this is done so that different transmitter occupies<br />

-11-


Analog Electronics /Introduction<br />

different frequency ranges. Then a receiver can separate the desired signal from the others by the<br />

use of frequency-selective electrical circuits called filters.<br />

When we approach the design of an electronic circuit to process a signal, one of our first questions<br />

should be "What is the frequency range of the signal" For example, we will see that ICs known as<br />

operational amplifiers can be very useful but they are limited to fairly low frequencies, usually below<br />

1MHz. If we need an amplifier for Channel 10 television we can rule out the use of operational<br />

amplifiers. Table 1.2 provides examples of the frequency ranges of some signals of practical interest.<br />

We conclude this section by noting that a signal can be represented either by the manner its<br />

waveform varies with time or in terms of its frequency spectrum. The two alternative<br />

representations are known as the time-domain and the frequency-domain representations,<br />

respectively. The frequency representation of a signal v(t) will be denoted by the symbol V(o>).<br />

TABLE 1.2<br />

FREQUENCY RANGES OF SELECTED SIGNALS<br />

SIGNAL<br />

Electrocardiogram<br />

Audible sounds<br />

Video signals<br />

AM radio broadcasting<br />

TV broadcasting<br />

FM radio broadcasting<br />

GSM cellular phone<br />

Satellite TV broadcasting<br />

FREQUENCY RANGE<br />

0.05-100Hz<br />

20 Hz-20kHz<br />

0-6 MHz<br />

150- 285kHz (LW)<br />

525 - 1600kHz (MW)<br />

48.5 - 56.5MHz (channel 1 VHF)<br />

222 - 230MHz (channel 12 VHF)<br />

470 - 478 MHz (channel 21 UHF)<br />

790 - 798MHz (channel 61 UHF)<br />

66 - 73MHz, 87.5 - 108MHz<br />

890 - 915MHz, 935 - 960MHz<br />

11.7 - 12.5GHz<br />

1.2 Analog and Digital Signals<br />

The voltage waveforms depicted in Fig. 1.2 are analog signals. The name derives from the fact<br />

that such a signal is analogous to the physical signal that it represents. The analog signal is a<br />

continuous function of time. Its magnitude can take on any value. Electronic circuits that process<br />

such signals are known as analog circuits. A variety of analog circuits will be studied in this<br />

subject.<br />

An alternative form of signal representation is that of sequence of numbers, each number<br />

representing the signal magnitude at an instant of time. The resulting signal is called a discrete-time<br />

signal or, shortly, discrete signal The signal is converted from analog to discrete form in the<br />

process of sampling. This process is illustrated in Fig. 1.4. The time instants t v /„ /,... are marked<br />

along the time axis. The magnitude of the signal is measured (sampled) at each of these time<br />

instants.<br />

Yet another useful form of signal representation is that of digital signal. It is obtained from a<br />

discrete signal by encoding the magnitude of signal samples with the use of finite-word-length<br />

binary code. In the example shown in Fig. 1.4, each sample value is represented by a 3-bit code<br />

word, corresponding to the amplitude zone into which the sample falls. Each sample value is<br />

converted into a code word, which in turn can be represented by a digital two-level waveform as<br />

shown in the lower panel of the figure. The digital signal is thus a sequence of encoded numbers.<br />

-12-


Analog Electronics /Introduction<br />

The circuit for conversion of signals in this manner is called an analog-to-digital converter<br />

(ADC). Conversely, a digital-to-analog converter (DAQ converts digital signals back to analog<br />

form. Now if we represent the magnitude of each of the signal samples in Fig. 1.4 by a number<br />

having a finite number of digits, then the signal magnitude will no longer be continuous; rather, it is<br />

said to be quantized or digitized. The resulting digital signal then is simply a sequence of gn^n^iypH<br />

numbers that represents the magnitudes of the successive signal samples.<br />

An example of analog, discrete and digital signal for a real world ECG waveform is shown in Fig.<br />

1.5. Please note that the discrete signal of Fig. 1.5(b) is defined only at the sampling instants. It no<br />

longer is a continuous function of time. However, since the magnitude of each sample can take any<br />

value in a continuous range, a discrete signal is still an analog signal. Perfect (error-free)<br />

reconstruction of analog signal is possible given the discrete-time signal, provided it has been<br />

sampled at sufficiendy high rate. Obviously, there is no way to recover the original analog signal<br />

from the digital signal. We will discuss these issues in more detail below.<br />

1.5<br />

1<br />

0.5<br />

0<br />

vffltmV]<br />

*5,<br />

0 0.1 0.2 0.3 0.4<br />

tim*[s)<br />

0.5 0.6 0.7 0.8<br />

'l.5<br />

v(n) [mV]<br />

10 20 30 40 50 60 70 80<br />

sampling instance indax, n<br />

Figure 1.5 Sampling of continuous-time analog signal (upper panel) results in the discrete-time signal (lower<br />

panel). Quantization of the discrete signal produces the following sequence of numbers:<br />

{0,0,5,11,14,14,11,7,4,1,0,1,1,1,2,3,3,3,3^6^0,51,90,105,77,27,-12,-28,-24,-10,<br />

-2,1,1,1,0,0,0,0,0,0,1,2,3,4,5,7,8,10,11,12,12,14,18^2^6,31,36,41,46,51,55,56,57,57,<br />

56,54,49,43,36,29,23,17,12,8,5,2,1,0,0} which represents the digital signal.<br />

The rate at which a signal must be sampled depends on the frequency content of the signal If a<br />

signal contains no components with frequencies higher than fy, the signal can be exacuy<br />

-13-


Analog Electronics /Introduction<br />

reconstructed from its samples, provided that the sampling rate is selected to be more than twice f b .<br />

(This is known as Shannon-Kotelnikov theorem.) For example, audio signals have a highest<br />

frequency of less than 20kHz. Therefore, the minimum sampling rate that should be used for audio<br />

signals is 40 kHz. Practical considerations require selection of a sampling frequency somewhat<br />

higher than the theoretical minimum. For instance, audio compact-disc technology converts audio<br />

signals to digital form with a sampling rate of 44.1kHz. Naturally, it is desirable to use the lowest<br />

practical sampling rate to minimize the amount of data (in the form of code words) to be stored,<br />

transmitted or manipulated.<br />

The second consideration important in converting analog signals to digital form is the number of<br />

magnitude zones to be used. Exact signal amplitudes cannot be represented, because all amplitudes<br />

falling into a given zone have the same code word. Thus, when a DAC converts the code words to<br />

form the original analog waveform, it is only possible to reconstruct an approximation to the<br />

original signal - the reconstructed voltage is constant within each zone. This is illustrated in Fig. 1.6.<br />

Thus some quantization error exists between the original signals and the reconstruction. Using a<br />

large number of zones can reduce this error, which requires a longer code word for each sample<br />

The number N of amplitude zones is related to the number of bits k by<br />

N = 2* (1.1)<br />

Thus if we are using an 8-bit {k=S) ADC, there are N=f—256 amplitude zones. In compact-disc<br />

technology, 16-bit words are used to represent sample values. With this number of bits, it is rather<br />

difficult for a listener to detect the effects of quantization error on the reconstructed audio signal.<br />

Reconstruction<br />

CMfclMtftiS I .—Analog signal<br />

jj I \ 1^ I Quantisation<br />

" *^I/ / *" ~ error<br />

,p - f- - t- - i Y<br />

Figure 1.6 Quantization error occurs when analog signal is reconstructed from its digital form<br />

Electronic circuits that process digital signals are called digital circuits. The digital computer is a<br />

system constructed mostly of digital circuits. Digital processing of signals has become quite popular<br />

primarily because of the tremendous advances made in the design and fabrication of digital circuits.<br />

Digital processing of signals is economic and reliable. Furthermore it allows a wide variety of<br />

processing functions to be performed - functions that are either impossible or impractical to<br />

implement by analog means. Nevertheless, most of the signals in the physical wodd are analog<br />

Also, there remain many signal-processing tasks that are best performed by analog circuits. They<br />

refer especially to high-frequency signals, but apply even to digital circuits themselves, which in fact<br />

are analog dynamical systems that we interpret as digital ones. It follows that a good electronics<br />

engineer must be proficient in both forms of signal processing. Such is the philosophy adopted in<br />

the contemporary electronic engineering study programs.<br />

The advantages of digital electronics are either coming through the use of digital electronics or are<br />

related to limitations of analog electronics. They are listed below.<br />

(1) Advantages of digital signal processing that originate in the features of digital electronic systems<br />

• flexibility (digital circuits are programmable),<br />

• lower sensitivity to external (e.g. temperature) and internal (e.g. aging and drift) effects,<br />

-14-


Analog Electronics /Introduction<br />

• accuracy can be controlled by selecting the word length to represent signal samples,<br />

• circuits are reproducible (no trimming or tuning during manufacture),<br />

• circuits are easier to manufacture in IC technology (no large Ls or Cs).<br />

(2) Advantages of digital signal processing related to limitations of analog electronics<br />

• "ideal memory" to store signals for an infinite time is possible with the digital techniques;<br />

thus very low frequency signals can be processed with no need for .large Ls and Cs,<br />

• linear phase filters - not available in analog electronics,<br />

• circuits for exact compensation of two effects,<br />

• adaptive systems,<br />

• precise signal transforms,<br />

• possibility of processing 2D signals (images).<br />

The main disadvantages of digital electronics are as follows:<br />

• more supply power required (passive digital circuits do not yet exist),<br />

• restricted to low-frequency applications,<br />

• when used in analog environment, often complex AD and DA converters are required<br />

• difficulties with AD and DA conversion of very weak and very strong signals;<br />

complicated analog pre- and post-amplifiers are required in such cases,<br />

• the same information (e.g. music signal) requires larger bandwidth as a digital signal than<br />

it does as the analog one.<br />

The analog circuits are<br />

• less accurate,<br />

• sensitive to noise,<br />

• advantageous for high-frequency, small- and large-signal applications.<br />

Modem and future systems contain both analog and digital circuits. We call them mixed-sigi* «J<br />

systems.<br />

1.3 Electronic System Block Diagrams<br />

Electronic systems are composed of subsystems or functional blocks. These functional blocks can<br />

be categorized as amplifiers, filters, signal sources, wave-shaping circuits, digital logic<br />

circuits, power supplies and converters. Briefly, we can say that amplifiers increase the power<br />

level of signals, filters separate desired signals from undesired signals and noise, signal sources<br />

generate waveforms such as sinusoids or square waves, wave-shaping circuits change one waveform<br />

into anomer (sinusoid to square wave, for example), power supplies provide necessary dc (direct<br />

current) power to the other functional blocks, and converters change signals from analog form to<br />

digital form, or vice versa. In section 1.6 we will consider the external characteristics of amplifiers in<br />

some detail.<br />

The block diagram of a typical AM radio is shown in Fig. 1.7, as an example of a simple electronic<br />

system. Notice that mere are three amplifiers and two filters. The local oscillator is an example of a<br />

signal source, and a peak detector is a special case of wave-shaping circuit The complete system<br />

description would include detailed specification of each block. For example, the gain, input<br />

impedance, output impedance, the bandwidth (range of frequencies where the gain does not drop<br />

below a specified value) of each amplifier would be given. (We define these terms later in uiis<br />

chapter.) Each functional block in turn consists of a circuit composed of resistors, capacitors,<br />

transistors, integrated circuits, inductive coils, and other devices.<br />

-15-


Analog Electronics /Introduction<br />

The main goal of this text is to introduce basic electronic circuits comprising diodes, bipolar<br />

transistors, field-effect transistors and opamps, in terms of their function and principle of operation.<br />

At present, due to the ever-increasing complexity of electronic circuits and systems, no circuit can<br />

be designed properly without the use of computers and adequate CAD software. We will then use<br />

one of the most popular programs for circuit simulation, called PSpice, to investigate the properties<br />

of the circuits under study. This approach, integrating the theory and simplified mathematical<br />

analysis with computer laboratory exercises will, we hope, prepare the readers for the circuit design<br />

methodologies they will "encounter as practicing engineers. A set of practical laboratory exercises is<br />

provided also to further strengthen the knowledge and to familiarize with measurement techniques<br />

typical for low-frequency analog electronics.<br />

1.4 Information-Processing Electronics versus Power Electronics<br />

Many electronic systems fall into one or more of the following categories: communication systems,<br />

medical electronics, instrumentation, control systems, and computer systems. A unifying aspect of<br />

these categories is that diey all involve collection and processing of information-bearing signals.<br />

Thus the primary concern of many electronic systems is to extract, store, transport or process<br />

information in a signal.<br />

Often, systems are also required to deliver substantial power to an output device. Certainly, this is<br />

true in an audio system for which power must be delivered to a speaker to produce the desired<br />

sound level A cardiac pacemaker uses information extracted from the electrical signals produced by<br />

the heart to determine when to apply a stimulus in the form of a minute pulse of electricity to<br />

ensure proper pumping action. Although the output power of a pacemaker is very small, it is<br />

necessary to consider the efficiency of its circuits to ensure long battery life.<br />

Many systems are concerned mainly with the power content of signals rather than information. For<br />

example, we might want to deliver ac (alternating current) electrical power, converted from dc<br />

supplied by batteries, to a computer even when the ac line power fails.<br />

1.5 Behavior of Circuit Components versus Frequency<br />

The way that a component behaves and the theoretical model we must use to represent it depend<br />

on the frequency of operation. For example, a simple 1000Q resistance can model the 1000Q-<br />

0.25W carbon-film resistor at frequencies of a few kilohertz, but at several hundred megahertz the<br />

more complex circuit model shown in Fig. 1.8 is needed for good accuracy. The small capacitance<br />

in parallel with the resistance has such a high impedance at low frequencies that it can be neglected.<br />

-16-


Analog Electronics /Introduction<br />

However, it cannot be neglected at high frequencies. Similarly, the inductance has very low<br />

impedance and can be neglected at low frequencies. (The inductance is associated with the magnetic<br />

field surrounding the leads when current flows in the resistor. Thus the exact inductance value<br />

depends on die length of the wire leads used in making connection to the remainder of the circuit.)<br />

We see later that the circuit models used for transistors at high frequencies include capacitances that<br />

can be neglected at low frequencies.<br />

•A/W<br />

1000 ohm<br />

w<br />

nH<br />

Figure 1.8 Circuit model for a 1000ft 0.25W carbon-film resistor<br />

Even the construction methods for a circuit depend on its operating frequency range. Many circuits<br />

intended to operate at low frequencies, say below 100 kHz, can be constructed by plugging<br />

individual components into a prototype board and wiring them together without much regard to<br />

the length of connecting wires or the distance between them. On the other hand, circuits intended<br />

to operate at several hundred megahertz must be carefully constructed with regard to the layout and<br />

lead length.<br />

It is somewhat difficult to make a definite boundary between "low" frequencies for which stray<br />

inductance and capacitance are not troublesome and "high" frequencies, for which they must be<br />

considered. In a high-impedance circuit for which the components have impedance of several<br />

megaohms or more, stray capacitance will be significant at lower frequencies than for a lowerimpedance<br />

circuits with impedance less than 100Q. In general, however, we do not need to<br />

consider stray wiring effects below about 100kHz. Usually, we must take them into consideration<br />

above about 10MHz.<br />

1.6 Summary<br />

Main function of electronic circuits is to process or generate signals. Either the Thevenin form or<br />

the Norton form can represent an electrical signal source. The sine-wave signal is completely<br />

characterized by its peak value (or rms value which is the peak/ V2 ), its frequency, and its phase<br />

with respect to an arbitrary reference time. A signal can be represented either by its waveform<br />

versus time, or as a sum of sinusoids. The latter representation is known as the frequency spectrum<br />

of the signal. Analog signals have magnitudes that can assume any value. Electronic circuits that<br />

process analog signals are called analog circuits. Sampling the magnitude of an analog signal at<br />

discrete instants of time and representing each signal sample by a number, results in a digital signal.<br />

Digital signals are processed by digital circuits. Modem and future systems contain both analog and<br />

digital circuits. Systems are composed of subsystems or functional blocks. The functional blocks are<br />

signal sources, amplifiers, niters, wave-shaping circuits, digital logic circuits, power supplies and<br />

converters. Electronic systems fall into the category of either information-processing or powerprocessing<br />

electronics. Any electronic system is made up of components (resistors, capacitors,<br />

inductors, semiconductor devices, etc.). Electrical properties of components depend on the signal<br />

frequency.<br />

BIBLIOTEKA GtOWNA PL<br />

W36^ S<br />

I<br />

-17-


Analog Electronics /Amplifiers<br />

2 Amplifiers: Behavioral Description<br />

In this chapter we introduce most important functional blocks that are employed in almost every<br />

electronic system, namely signal amplifiers. We will define specifications and characteristics of<br />

amplifiers, which describe their behavior as seen by an external observer. In this system-like<br />

approach we will not discuss internal structure of amplifiers, leaving this topic for the later chapters.<br />

2.1 Basic Concepts<br />

Signal amplification is conceptually the simplest signal-processing task. The need for amplification<br />

arises because transducers provide signals that are said to be "weak", that is in the microvolt (uV)<br />

and millivolt (mV) range. An example is a signal from a microphone, which is of about lmV peak<br />

as one speaks to this transducer. Such a weak signal cannot produce any noticeable acoustic effect if<br />

used to drive a loudspeaker. Much stronger signal is needed for a loudspeaker in order to obtain a<br />

louder version of the sound entering the microphone. Thus the signal from the microphone is used<br />

as the input to an amplifier with a voltage gain of 10000 to produce an output signal which is a peak<br />

value of 10 V. This "strong" signal is applied to the loudspeaker which produces a loud replica of<br />

the microphone's input.<br />

The concept of signal amplification is illustrated in Fig. 2.1. The signal source produces a signal t\(t)<br />

that is applied to the input terminals of the amplifier, which generates an output signal<br />

v o (0 = 4,v,(0 (2-1)<br />

across the load resistance R L connected to the output terminals. The constant A^ is called the<br />

voltage gain of the amplifier. Often, the voltage gain is much larger than unity, but we will see later<br />

that useful amplification can take place even if the magnitude of A, is less than unity.<br />

Sometimes A^ is a negative number, so the output voltage is an inverted version of the input, and<br />

he amplifier is then called an inverting amplifier. On the other hand, if A^ is a positive number,<br />

we have a noninverting amplifier. These notions are illustrated in Fig. 2.2.<br />

Equation (2.1) is a linear relationship; hence the amplifier it describes is a linear amplifier. A linear<br />

amplifier does not introduce any distortions to sinusoidal amplified signals. The output signal is an<br />

exact replica of the input, except of course for having larger amplitude (and possibly a nonzero<br />

phase shift with respect to the input, which is discussed in Section 2.6). Linearity is the muchneeded<br />

feature of amplifiers so that the information contained in die signal is not changed and no<br />

new information is introduced. Any change in waveform is considered distortion and is obviously<br />

undesirable.<br />

-18-


Analog Electronics/Amplifiers<br />

Figure 2.2 Input signal to an amplifier (upper panel), output signal of a noninverting amplifier of gain A, =8,<br />

output signal of an inverting amplifier of gain A y =-8\<br />

The signal amplifier is a two-port network. This is clearly seen in Fig. 2.1 where the two input<br />

terminals are distinct to the two output terminals. A more common situation however is illustrated<br />

in Fig. 2.3 where a common terminal exists between the input and the output The common<br />

terminal is used as a reference point and is called the circuit ground.<br />

Figure 2.3 An amplifier with a common terminal (ground) between the input and the output ports.<br />

Voltage amplification can be modeled as a voltage controlled voltage source as illustrated in Fig. 2.4.<br />

Because real amplifiers draw some current from the signal source, a realistic model of an amplifier<br />

must include a resistance r\ across the input terminals. Furthermore, a resistance R^ must be<br />

included in series with the output terminals to account for the fact that the output voltage of an<br />

amplifier is reduced when load current flows.<br />

The input resistance R; of the amplifier is the equivalent resistance seen when looking into the<br />

input terminals. As we will find later, the input circuitry can sometimes include capadttve and<br />

inductive effects, and we would then refer to the input impedance. For example, the input<br />

amplifier of a typical oscilloscope has input impedance containing a 1-Mfi resistance in parallel with<br />

a 47-pF capadtance.<br />

Figure Z4 Model of an electronic amplifier<br />

The voltage-controlled voltage source in Fig. 2.4 models the amplification properties of the<br />

amplifier. Notice that the voltage produced by this source is the input voltage times a constant A°.<br />

-19-


Analog Electronics /Amplifiers<br />

If the load is an open circuit, R^—°o, there is no drop across the output resistance R^ then v Q =<br />

A%v t . For this reason the constant A% is called the open-circuit voltage gain. To summarize, the<br />

voltage-amplifier model has three parameters: the input impedance, the output impedance and the<br />

open-circuit voltage gain.<br />

As shown in Fig. 2.4, the input current i t is the current delivered to the input terminals of the<br />

amplifier and the output current to is the current flowing through the load. The cuttent gain A i of<br />

the amplifier is the ratio of the output current to the input current<br />

For linear amplifiers, the input current can be expressed as the input voltage divided by the input<br />

resistance. For linear load, the output current is the output voltage divided by the load resistance.<br />

(Linear amplifiers will be investigated in this chapter unless stated otherwise.) Thus we can find the<br />

current gain in terms of the voltage gain and the resistance as<br />

in which<br />

A= - (2-4)<br />

is the voltage gain with the load resistance connected. Usually, A^ is smaller in magnitude than the<br />

open-circuit voltage gain A°, because of the voltage drop across the output resistance.<br />

The power delivered to the input terminals by the signal source is called the input power P^ and the<br />

power delivered to the load is the output power P Q . The power gain G of an amplifier is the ratio<br />

of the output power to the input power<br />

G = £ (2.5)<br />

Because we are assuming that the input impedance and the load are purely resistive, the average<br />

power at either set of terminals is simply the product of die root-mean-square (rms) current and the<br />

rms voltage. Thus we can write<br />

Notice that we have used uppercase symbols, such as V 0 and J 0 , for the rms values of the currents<br />

and voltages. We use lowercase symbols, such as v 0 and i^ for the instantaneous values. Of course,<br />

since we have assumed that the instantaneous output is a constant times the instantaneous input,<br />

the ratio of the rms voltages is the same as the ratio of the instantaneous voltages, and both are<br />

equal to the voltage gain of the amplifier.<br />

Exetcise 2.1 An amplifier has an input resistance l\ = 2000 Q, an output resistance of 25 Q, and<br />

an open-circuit voltage gain of 500. The source has an internal voltage of V s =20mV and a<br />

resistance R t =500Q. The load resistance is R L =750. Find the voltage gains A y = VJV\ and A^=-<br />

VJ V s , the current gain and the power gain. Ans. ^=375,^=300,^=10 4 , G=3.75-10 6 .<br />

Exetcise 2.2 Assume that we can change the load resistance in Exercise 2.1. What value of load<br />

resistance maximizes the power gain What is the power gain for this resistance Ans. R L =250,<br />

G=5-10 6 .<br />

-20-


Analog Electronics /Amplifiers<br />

There exist other models of amplifiers, alternative to the voltage model. The input and output<br />

resistances are the same for all models. They differ by the type of controlled source at the output.<br />

Current amplifier model employs a current controlled current source, transconductance<br />

amplifier uses a voltage-controlled current source and the transresistance amplifier is the one<br />

with current-controlled voltage source at the output. The choice of the particular model is the<br />

matter of analytical convenience or feasibility of measurement of the model parameters. An<br />

amplifier can be modeled by any of the four models, provided that neither of the resistances (input<br />

or output) is zero or infinity. For example, if JRj^O, then t»,=0 and the voltage gain A^— v 0 /i\ is not<br />

defined.<br />

One can easily show that if the input impedance of an amplifier is much higher than the internal<br />

impedance of the source, the voltage produced across the input terminals is-nearly the same as the<br />

internal source voltage. On the other hand, if the input impedance is very low, the input current is<br />

neady equal to the short-circuit current of the source. From this point of view, voltage amplifiers<br />

should be designed to have large input impedance and current amplifiers should have very low<br />

input impedance. Taking the output impedance level into consideration, we can force a desired<br />

voltage waveform to appear across a variable load by designing the amplifier to have very low<br />

output impedance compared to the load impedance. An example is an audio amplifier with a<br />

variable number of loudspeakers connected to it On the other hand, we can force a given current<br />

waveform through a variable load by designing an amplifier to have very high output impedance<br />

compared to the load impedance. Here an example could be the amplifier in optical communication<br />

system where a light-emitting diode (LED). It is used to produce a light wave whose intensity is<br />

proportional to a message signal such as a voice waveform. Because LED has nonlinear relationship<br />

between voltage and current, light intensity "is not" proportional to the voltage across this device,<br />

whereas it is proportional to the current flowing through the diode. Thus LED should be driven<br />

from a current amplifier in this example.<br />

We see that certain applications call for amplifiers with very high or very low input impedance<br />

(compared to the source) and very high or very low output impedance (compared to the load). Such<br />

amplifiers are classified as follows.<br />

An ideal voltage amplifier (voltage-controlled voltage source) senses the open-circuit voltage of<br />

die source and produces and amplified voltage across the load - independent of the load impedance.<br />

Thus the ideal voltage amplifier has infinite input impedance and zero output impedance.<br />

Figure 26 An ideal current amplifier (CCCS)<br />

-21-


Analog Electronics /AmpMers<br />

An ideal current amplifier (current-controlled current source) senses the short-circuit current of<br />

the source and forces an amplified version of this current to flow through the load. Thus an ideal<br />

current amplifier has zero input impedance and infinite output impedance.<br />

O 4 O<br />

Figure 2.7 An ideal ttansconductanceamplifier (VCCS)<br />

An ideal transconductance amplifier (voltage-controlled current source) senses the open-circuit<br />

voltage and forces a current proportional to this voltage to flow through the load. Thus it has<br />

infinite input impedance and infinite output impedance.<br />

Figure 2.8 An ideal transresistance amplifier (CCVS)<br />

An ideal transresistance amplifier (current-controlled voltage source) senses the short-circuit<br />

current of the source and produces a voltage proportional to this current to appear across the load.<br />

Thus the ideal transresistance amplifier has a zero input impedance and zero output impedance.<br />

The amplifier models considered here are unilateral; that is signal flow is unidirectional, from input<br />

to output Most real amplifiers show some reverse transmission, which is usually undesirable but<br />

must nonetheless be modeled. We will return to this point later.<br />

Recall that we originally defined the gain of an amplifier to be the ratio of the output signal to the<br />

input signal (2.4). This is true for amplifiers that do not contain any energy-storing element.<br />

However, if any time delay or linear distortion occur, the ratio of output to input is a function of<br />

time rather than a constant Thus we should not try to find the gain of an amplifier by taking the<br />

ratio of the instantaneous output to input. Instead we recognize that gain is a function of frequency<br />

and take the ratio of phasors for a sinusoidal input signal to find the (complex) gain at each<br />

frequency<br />

2.2 Cascaded Amplifiers<br />

Amplifiers are built using active devices - most often transistors. We will see later that the gain<br />

obtainable from a single-transistor amplifier is limited. Thus there is often a need to combine a<br />

number of amplifiers in order to amplify a weak signal to the desired level. This can be achieved by<br />

connecting the output of one amplifier to the input of another, as shown in Fig. 2.9. This is called a<br />

cascade connection of-the amplifiers. The overall voltage gain of the cascade connection is given<br />

by<br />

A v = ^ (2.7)<br />

Multiplied and divided by f ol this becomes<br />

-22-


Analog Electronics /Amplifiers<br />

" v n v 0l<br />

However, referring to Fig. 2.5, we see that v^v oV<br />

Therefore we can write<br />

(2.8^<br />

'<br />

v ,i<br />

v ,2<br />

We note also ^3OSAA^X=V OX / » A is the gain of the first stage andy4 v2 =f o2 / tfo is the gain of the second<br />

stage, so we have<br />

A v =A vl A v2 (2.10)<br />

Thus the overall gain of cascaded voltage amplifiers is the product of the voltage gains of the<br />

individual stages. (Of course, it is necessary to include loading effects in computing the gain of each<br />

stage. Notice that the input resistance of the second stage loads the first stage.) Similar rules can be<br />

derived for the number of cascaded amplifiers bigger than two can.<br />

(2.9)<br />

Exercise 2 J Show that the overall current gain of a cascade connection of amplifiers is the product<br />

of the current gains of the individual stages and that the overall power gain is the product of<br />

individual power gains.<br />

2.3 Power Supplies and Efficiency<br />

For appropriate operation of the internal circuitry of an amplifier, power must be supplied to it<br />

from an external power supply. The power supply typically delivers current from several dc<br />

voltages, see an example in Fig. 2.10. The average power supplied to the amplifier by each voltage<br />

source is the product of the average current and voltage. The total power supplied is the sum of the<br />

powers delivered by each source. For example, the total average power supplied to the amplifier of<br />

Fig. 2.10 is<br />

*s - 'AA^A + *BB*B<br />

(2.11)<br />

We have seen in Exercise 2.1 that the power gain of an amplifier can be very large. Thus the output<br />

power delivered to the load is much greater than the power taken from the signal source. This<br />

additional power is taken from the power supply. For example, a stereo audio system converts part<br />

of the power taken from the power supply into signal power that is finally converted to sound by<br />

the loudspeakers. Part of the supplied power is also dissipated as heat in the internal circuits of the<br />

amplifier. This dissipation is an undesirable effect since the dissipated power is lost in most cases<br />

and makes the battery discharged too fast in case of mobile battery-powered devices. We usually try<br />

to minimize this effect when designing the internal circuitry of amplifiers.<br />

-23-


_ Analog Electronics /Amplifiers<br />

Figure 2.10 TTie power supply delivers power to the amplifier from<br />

several external constant voltage supplies.<br />

From energy conservatiqn law, the sum of the power entering the amplifier from the signal<br />

P { and the power from the power supply P s must be equal to the sum of the output power<br />

the power dissipated P d .<br />

p i+ p s = p 0 +p d (2.12)<br />

:>wer is negligible compared to the other terms in this<br />

percentage of the power supplied that is converted in<br />

P.<br />

n =<br />

P,+P s<br />

(2.13)<br />

The power and efficiency values given in Exercise 2.4 below are typical of one channel of a stereo<br />

amplifier under high output test conditions. Maximi2ing the efficiency is one of the important<br />

design goals for battery-operated devices such mobile phones, hearing aids, implantable heart<br />

pacemakers, etc., as well as for high-power equipment, such as motor drivers, power regulators,<br />

DC-DC converters, etc.<br />

Exercise 2.4 Find the input, output, supply and dissipated power in the amplifier shown in Fig. 2.6,<br />

assuming the following values of its parameters: v s = 1 mV^, J^ = 0, ^ = 100 kQ, A% = 10 4 , R^ =<br />

2Q,R L =8Q, Vj^ = 15 V, J A = 1 A, V m = 15 V, J B = 0.5 A. Calculate the efficiency of the<br />

amplifier. Am. P k = 10 pW, P 0 = 8 W, P s = 22.5 W, P d = 14.5 W, rj = 35.6%.<br />

Most amplifiers discussed in this lecture notes are small-signal circuits. The power involved is very<br />

small and the efficiency was not a concern. We do not consider power amplifiers, or output<br />

stages that are expected to provide large signal power requited by their loads. Examples of loads<br />

for power amplifiers are stereo and public address loudspeakers, deflection coils in video monitors,<br />

and servomotors in X-Y plotters. Powers supplies are also good examples of power circuits; their<br />

function is to provide the dc power required by digital and analog components of all lands.<br />

Several special concerns preoccupy the designer of an amplifier that must deliver a large amount of<br />

power to the load. One is the amplifier efficiency, as discussed above in this Section. Equation<br />

(2.13) shows that, for given output power, low efficiency directly means greater demand for<br />

supplied power. Furthermore, efficiency suggests a second issue - the ability of the circuit<br />

components to dissipate heat Any power that does not leave the circuit as the load power<br />

contributes to the heating of transistors and resistors. In power circuits special efforts should be<br />

-24-


Analog Electronics /Amplifiers<br />

made to ensure that components are not desttoyed by excessive heat. Finally, the large signal<br />

amplitudes necessary for large output signal power make nonlinear distortion (see Section 2.9) a<br />

matter of concern. Small signal models, based upon linearization of the device characteristic at some<br />

jj-point (Sections 3.4, 4.6, 5.7 and 5.11), no longer apply here. Large-signal dynamic models have to<br />

be used for circuit analysis and design, such as SPICE models discussed in Sections 3.8, 4.10 and<br />

5.14.<br />

Power amplifiers are classified according to the fraction of the time an output power transistor<br />

conducts the current during one signal cycle. Class A amplifiers have output transistors in which<br />

signal current is nonzero all the time. The FET amplifiers discussed in Sections 4.7 and 5.13 are all<br />

class A amplifiers. The dc current of a class A amplifier is the same, no matter the signal is zero or<br />

nonzero. For greater efficiency, class B amplifiers employ transistors that are active only half time -<br />

otherwise they are cut-off. At zero signal value, there is no dc current through- a class B output stage<br />

- and no power consumption. An example of PSpice simulation of a class B bipolar transistor<br />

amplifier is presented in Section 6.1.3. So-called crossover distortion appears in class B amplifiers<br />

(see Section 6.1.3), so for their reduction class AB amplifiers are designed. In class AB circuits,<br />

transistors conduct slightly more than half the signal cycle. Class AB circuits have efficiencies<br />

approximately the same as for class B circuits, but produce less distortion.<br />

Class C amplifiers deliver large amount of output power at high efficiency by employing an output<br />

transistor that conducts output current for only small fraction of a cycle. The resulting short,<br />

periodic pulses of output current excite a resonant circuit, which suppresses the distortion<br />

components that arise from the nonlinear operation of the transistor.<br />

Class D amplifiers produce binary output waveforms of very high power with efficiencies<br />

approaching 100% by using transistors as switches. In analog applications, the class D amplifier<br />

includes a modulator that first transfers information from waveform amplitude into a more suitable<br />

form as pulse width. A lossless filter than removes undesired high-frequency terms from the output.<br />

Applications of power amplifiers, especially class C and D circuits, are somewhat specialized.<br />

Therefore, they are not discussed further in this text<br />

2.4 Decibel Notation<br />

As we have noticed from the Exercises, the gain of an amplifier can take on values spanning a very<br />

large range of magnitudes. To facilitate visualization of these extreme values, amplifier gain is often<br />

expressed with a logarithmic measure. Specifically, the voltage gain A^ can be expressed in decibels<br />

(dB)as<br />

^=201og|^v|dB (2.14)<br />

and the current gain A, can be expressed as<br />

^=201ogM,|dB (2.15)<br />

Since power is related to voltage (or current) squared, the power gain G can be expressed in<br />

decibels as follows<br />

G^g^OlogG dB (216)<br />

A power gain of G=100 converts to 20dB, unky gain converts to OdB, and so on. The absolute<br />

values of the voltage and current gains are used because in the case of inverting amplifiers A^oxA^<br />

can be negative values. A negative gain A v means that there is a 180°-phase difference between die<br />

input and output signals; it does not imply mat the amplifier attemiates the signal On the other<br />

-25-


AmJofElectronica/AinptiSen<br />

hand, an amplifier whose voltage gain is -20 dB is, in fact attenuating the input signal by a factor of<br />

10 (that is ^=0.1).<br />

Recall that the overall gain for cascaded amplifiers is the product of the power gains of the<br />

individual amplifiers. Wbeti fa gm an e>pnssedmdecfals, the gons of cascaded stagum<br />

the properties of the logarithm function. To illustrate this point we have<br />

G = G X G 2 (2.17)<br />

When expressed in decibels, this becomes<br />

G* = lOlogG - lOlogCG.Gj) = lOlogCG,)* 101og(G 2 ) dB (2.18)<br />

Note, referring back to (2.6), that power gain in decibels is equal to voltage gain in decibels only<br />

when RL=R}.<br />

Electronics engineers often use decibel notation for voltages, currents, powers, or other quantities.<br />

To do so, a reference level must be stated or implied The quantity to be expressed in decibels is<br />

divided by the reference value and the ratio is converted to decibels by taking 20 times me<br />

logarithm of the ratio for voltages or currents. Ten times the logarithm of the ratio is taken for<br />

powers. Some commonly used reference levels are 1 volt (dBV), 1 mW (dBm), and 1 watt (dBW).<br />

For example, 40 dBV is a designation of 100 V, -10 dBm is for 0.1 mW, -40 dBW is also 0.1 mW,<br />

and so on.<br />

2.5 Frequency Response<br />

From Chapter 1.1 we know that the input signal to an amplifier can always be expressed as the sum<br />

of sinusoidal signals. So far we have considered the gain parameter to be a constant However, since<br />

physical amplifiers contain capatiuve and inductive elements, the gain is a function of frequency.<br />

Furthermore, the amplifier affects the amplitude as well as the phase of the input signal<br />

It follows then that an important characterization of an amplifier is in terms of its response to input<br />

sinusoids of different frequencies. Such characterization of amplifier performance is known as the<br />

amplifier frequency response.<br />

26


Analog Ehcttoaka/AmpWka<br />

Figure 212 Measuring the frequencyresponse of a Knear amplifier (a), phase characteristic of a two-transistor<br />

small-signal amplifier (b) whose amplitude characteristic is shown in Fig. 2.7<br />

Figure 2.12a shows a linear voltage amplifier fed at its input with a sine-wave signal of amplitude V {<br />

and frequency CD. As the figure indicates, the signal measured at the amplifier output is also<br />

sinusoidal with exactly me same frequency 0). This is always true for linear circuits. However, the<br />

output sinusoid will have in general different amplitude and a different phase relative to the input.<br />

The ratio of the amplitude of me output sinusoid (VJ to the amplitude of the input sinusoid (V$ is<br />

the amplifier gain or transmission at the test frequency 0). Hie angle


Analog Electronics /Amplifiers<br />

as the midband region. This is typical for wideband amplifiers. Wideband amplifiers are used for<br />

signals that occupy a wide range of frequencies, such as audio signals or video signals (see Table<br />

1.1). In some cases, the frequency response of an amplifier is deliberately limited to a small<br />

bandwidth compared to the center midband frequency. Such an amplifier is called a nattowband<br />

or banpass amplifier. Bandpass amplifiers are frequency selective circuits used in radio receivers to<br />

amplify the signal from one transmitter and reject the signals from other transmitters in adjacent<br />

frequency ranges.<br />

Figure 2.13 Gain versus frequency for wideband amplifiers: ac-coupled amplifier (a) and<br />

dc-coupled amplifier (b)<br />

Usually we specify the approximate useful frequency range of an amplifier by giving the frequencies<br />

at which the voltage (or current where appropriate) gain magnitude is l/\2 times the midband gain<br />

magnitude. These are known as half-power frequencies (cut-off frequencies) because the output<br />

power level is half the value for the midband region if the constant-amplitude, variable-frequency<br />

sinewave test input is used. Expressing the factor 1 /"V2 in decibels results in -3.01 dB. Thus at the<br />

half-power frequencies, the voltage or current gain is approximately 3dB lower than the midband<br />

gain. The bandwidth B of an amplifier is the distance between the half-power frequencies, {f H -f{).<br />

This is illustrated in Fig. 2.13.<br />

In the cases illustrated by Fig. 2.13(a), the gain drops to zero at dc (zero frequency). Such amplifiers<br />

are called ac-coupled, because only ac signals are amplified. These amplifiers are often constructed<br />

by cascading a number of amplifiers that are connected together by coupling capacitors, so that<br />

dc voltages of signal source, and individual amplifier stages are separated from each other. This is<br />

illustrated in Fig. 2.14 where; caparitive coupling prevents a dc input component from affecting the<br />

first stage, dc voltages in the first stage from reaching the signal source and the second stage, and dc<br />

voltages in the second stage from reaching the first stage and the load. Sometimes magnetic<br />

transformers are used for coupling, which also leads to an ac-coupled amplifier with a zero gain at<br />

dc.<br />

-28-


Analog Electronics /Amplifiers<br />

Other amplifiers have constant gain all the way down to the dc, as shown in Fig. 2.13(b). They are<br />

said to be dc-coupled or direct coupled. Amplifiers that are realized as integrated circuits are<br />

often dc coupled because the capacitors or transformers needed for ac coupling cannot be<br />

fabricated in integrated form.<br />

Electrocardiograph amplifiers are deliberately ac-coupled because a dc voltage of nearly a volt often<br />

occurs in the input due to contact potentials developed on the electrodes placed on the skin. The ac<br />

signal generated by the heart is on the order of lmV, and therefore the gain of the amplifier is high -<br />

typically 1000 or more. A lV-dc input would cause the amplifier to produce an output of 1000V. It<br />

would be difficult and undesirable to design an amplifier capable of such large outputs. Therefore, it<br />

is necessary to ac-couple the input circuit of an electrocardiograph to prevent the dc component<br />

from overloading the amplifier.<br />

Amplifiers for video signals need to be dc coupled because video signals have frequency<br />

components from dc to about 6MHz, see Table 1.1. Dark pictures result in a different dc<br />

component than bright on the average pictures. It is necessary to dc-couple a video amplifier in<br />

order to preserve dc component and to obtain proper brightness of the image.<br />

As indicated in Fig. 2.13, the gain of an amplifier always drops off at high frequencies. This is<br />

caused either by small capacitance in parallel with the input or output terminals or by small<br />

inductances in series with the signal path in the amplifier circuitry. This is illustrated in Fig. 2.10.<br />

Recall that impedance of a capacitor is inversely proportional to frequency, resulting in effective<br />

short circuit at sufficiendy high frequencies. The impedance of an inductor is proportional to<br />

frequency, so it becomes an open circuit at very high frequencies.<br />

Figure 2.15 Capacitance in parallel with the signal path and inductance in series with the signal path reduce<br />

gain in the high-frequency region<br />

Some of these small capacitances occur because of stray wiring capacitance between signal-carrying<br />

conductors and ground. Other capacitances are integral parts of active devices (transistors)<br />

necessary for amplification. Small inductances result from the magnetic fields surrounding the<br />

conductors in the circuit For example, a critically placed piece of wire 1-cm long can have enough<br />

inductance to limit severely the frequency response of an amplifier intended to operate at several<br />

GHz.<br />

2y


Analog Electronic* /AupttiBen<br />

Fot frequencies up to several MH2, the circuit and device capacitances ate the main source of<br />

frequency response roll-off, either at low or high frequencies. To illustrate their impact on<br />

frequency response we will consider an amplifier that can be represented by a real transconductance<br />

amplifier model (Figure 2.16). Such a model is an ideal transconductance amplifier as shown in<br />

Figure 2.7, augmented with input nonzero resistance r, and output finite resistance r* It is an<br />

adequate low-frequency representative of amplifying devices, such as field-effect and bipolar<br />

transistors. For the field-effect transistor, the input resistance r t is infinite whereas for the bipolar<br />

transistor a base resistance is added in series with the input terminal as will be discussed in Chapters<br />

4 and 5, respectively.<br />

Figure 2.16 Low frequencynon-ideal transconductance amplifier model<br />

The input to the amplifier is voltage source it with internal resistance R». Resistance RL represents<br />

the load They are shown in Fig. 2.17. The symbol ^n used in Figs 2.16 and 2.17 to denote the shortcircuit<br />

transconductance coefficient is equivalent to A& in Fig. 2.7. To illustrate the effect of the<br />

source resistance on the frequency response, we will use a voltage gain definition slightly different<br />

to that given by Equation (2.1), namely<br />

4. = 7T (2- 21 )<br />

Figure 217 Transconductance amplifier<br />

There is no capacitive or inductive element in the schematic diagram of Fig. 2.17. Therefore, the<br />

bandwidth of this circuit is infinite. Indeed, the output voltage<br />

r 0 R<br />

V L<br />

0 =-ZmVi<br />

r 0 +R]L<br />

does not depend on frequency. The voltage gain of the amplifier in Figure 2.17 can be expressed as<br />

r 0Ri<br />

Ays ~ ~8lH "" Ay, (2.23)<br />

r >so<br />

0 +R L r { +R s<br />

The coefficient Ano in Equation (2.23) is a constant as plotted in Figure 2.18(a) wherein = 50 mS,<br />

R, = 600 Q, r, = 5 kO, r 0 = 100 kQ, and RL = 5.6kft were assumed. The phase characteristic in<br />

Figure 2.18(b) is also constant and equal to 180 degrees that means the amplifier under<br />

consideration is an inverting one. Of course, the circuit discussed is an idealized model of real-world<br />

amplifiers. Nevertheless, for medium frequency range it is an adequate model of transistor<br />

amplifiers.<br />

Low frequency effects. To illustrate the effect of coupling capacitances on the frequency response<br />

of amplifiers we consider the circuit of Figure 2.19. It contains a coupling capacitor C, in series with<br />

the input resistance r f The output voltage is described by Equation (2.22) and the input voltage can<br />

be expressed as<br />

-30-


Analog Ekctroak* /AnpUtten<br />

Vi=V t<br />

R s +r t +<br />

1<br />

Jar t C s<br />

= V,<br />

(2.24)<br />

l+JG>(R s +r t y: s<br />

JG>(R s +ri)C s<br />

* rt,+r, 1+ ./*L<br />

R s +r i l+ja>T L<br />

Figure 219 Ac-coupled ttansconductanceamplifier<br />

where<br />

U=(*S+ri)C s (2.25)<br />

is the low-frequency time constant of the amplifier. Combining Equations (2.22) and (2.26), one<br />

obtains the effective voltage gain of the amplifier in Fig. 2.19<br />

Ays ~ 8n<br />

n JQ>r L r 0 R L<br />

R s +r i l + ja>T L r 0 +R l<br />

(2.26)<br />

~~Ayso<br />

1+JOTL<br />

Equation (226) indicates that in this case the voltage gain is a single-pole function of frequency.<br />

The voltage gain magnitude<br />

Vl+(a>rJ 2 (2.27)<br />

-31-


Analog Electronics /Amplifiers<br />

tends to zero for frequencies approaching zero. Thus an ac-coupled amplifier does not provide a dc<br />

gain. This is shown in Figure 2.20a where C s = \ \x¥ was assumed with the other circuit parameters<br />

taking the same values as for Figure 2.18.<br />

The gain increases with frequency. For high frequencies, the gain magnitude is equal to A yM . Then<br />

one can say the ac-coupled amplifier acts as a highpass filter. At the lower cut-off frequency, jL, the<br />

It is easy to show that for a single-pole transfer function of the type given by Equation (2.26), the<br />

lower cut-off frequency is inversely proportional to the low-frequency time constant<br />

h=^— (2-28)<br />

Using Equation (2.26) one can find expression that describes the phase characteristic of the<br />

amplifier in Figure 2.19<br />

ZA\, S =270° -tg-\coT L )<br />

(2.29)<br />

.-li<br />

=-90 -tg- J (a>r L )<br />

The phase characteristic (2.29) of the ac-coupled amplifier, computed using PSpice program is<br />

plotted in Figure 2.20b.<br />

High frequency effects. Capacitances in parallel with the signal path (as well as capacitances<br />

connected between the output and input terminals of inverting amplifiers that will be discussed later<br />

on) make the amplifier gain decreasing with frequency. To illustrate this effect, capacitance C 0 is<br />

connected in parallel to the load resistance of the amplifier, as shown in Figure 2.21. The output<br />

voltage is now given as<br />

gain magnitude is equal to A\, so I-J2 . The same results (except for different expression to describe<br />

the low-frequency time constant) would be obtained if a coupling capacitor is connected in series<br />

with the load resistance RL-<br />

-32-


Analog Electronics /AmpliBets<br />

v Q = -g m Vi<br />

jcoC 0<br />

R 0 +—t—<br />

JoCo<br />

Ro<br />

= -g m Vi<br />

\+jaC 0 R 0<br />

(2.30)<br />

= n R ~Sm's 0<br />

ri+Rs \+jaC 0 R 0<br />

-_«T V 1-<br />

R o<br />

where<br />

T H ~<br />

R o C o<br />

is the high frequency time constant of the amplifier and<br />

_ r Q R L<br />

Rn =<br />

r 0 + R L<br />

(2.31)<br />

(2.32)<br />

Figure 2.21 A transconductance amplifier with parallel capacitance at the output<br />

Substituting Equation (2.23) into (2.31), dividing both sides by V s and making use of the definition<br />

(2.21), one obtains the voltage gain<br />

4,=-,<br />

K ° (2-33)<br />

that is a single-pole function of frequency. Its magnitude<br />

I4«(


Antdog Electronics /Amplifiers<br />

The phase characteristic (2.36) of the amplifier in Figure 2.21 with parallel capacitance C = 10 pF,<br />

computed using the PSpice program, is plotted in Figure 2.22b.<br />

2.6 The Miller Theorem<br />

It is often the case in transistor circuits that a capacitor is connected between the output and input<br />

ports of the amplifier. Such a capacitor may represent a base-collector junction capacitance of a<br />

bipolar device or drain-gate capacitance of FET or MOSFET field-effect devices. As an example, a<br />

simplified equivalent circuit of such an amplifier is shown in Figure 2.23, where g„ = 50 mS, R, =<br />

600 CI, r,, = 5 kQ, r 0 = 100 \£i, R^ = 5.6kQ and C F = 10 pF are assumed.<br />

Figure 2.23 A transconductance amplifier with a feedback capacitor<br />

The capacitive element connected between the input and output terminals causes difficulties in<br />

analysis since its presence leads to higher-order nodal equations. The difficulties can be overcome<br />

by applying Miller theorem to the circuit This theorem is a basis to one of approximate methods of<br />

transistor amplifier analysis. To derive it, we will consider a general amplifier with feedback<br />

impedance as shown in Figure 2.24a.<br />

y,*<br />

O"o<br />

v, o 1<br />

-OVo<br />

\ ><br />

(a)<br />

(b)<br />

Figure 2.24 Original amplifier with feedback impedance (a), equivalent amplifier with feedback element split<br />

into two parts (b)<br />

-34-


Analog Electronics /Amplifiers<br />

For the original circuit we have<br />

j _VJ-VQ JTi-AVi =Vj>1. If Z is resistive, the impedances Zml and Zm2 are also resistive, with Zmi being much<br />

smaller than Z and Zm2 being approximately equal to Z. If Z is a capacitive element, Z=l/')(oC, the<br />

impedances Zmi and Zm2 are also capacitive<br />

C ml =C(l-A) (2.43)<br />

C m2 =C l -^j (2.44)<br />

i<br />

Figure 2.25 Circuit equivalent to the amplifier of Fig. 2.23<br />

Figure 2.25 shows the equivalent circuit for the amplifier of Figure 2.23. The approximate values of<br />

die Miller capacitances are<br />

C ml =C F (l-A vo ) (2.45)<br />

C m2 =C F 1 -^- (2-46)<br />

-A vo<br />

where Am = Vo/V{. Now one can derive die formula for the voltage gain of die amplifier. The<br />

output voltage is given by<br />

-35-


Analog Electronics /AmpliGcra<br />

"o ~ Sm"i, .<br />

l + JG>T 2<br />

where Ro is defined by Equation (2.32) and the output circuit time constant is described :das<br />

T 2 = R 0 C„2<br />

(2.48)<br />

In turn, for the input voltage we have<br />

1<br />

(2.49)<br />

where<br />

l ~ R \ C m\<br />

and<br />

V/<br />

*1 = *s+n<br />

(2.47)<br />

(2.50)<br />

(2.51)<br />

Combining Equations (2.47), (2.49) and (2.23), one obtains the voltage gain<br />

A =_, __J L_<br />

l + JCDTi l + JG)T 2<br />

(2.52)<br />

which is a two-pole transfer function. The time constants are related to the angular firequendes<br />

CO X = — (2.53)<br />

fi> 2 =— (2.54)<br />

*2<br />

called break firequendes, since for CO exceeding a value of any of these firequendes the slope of the<br />

amplifier amplitude characteristic increases by 20 dB/decade. Using (2.53) and (2.54) one can<br />

rewrite Equation (2.52) as follows<br />

Ay S — ~A VS0<br />

CO<br />

•I'<br />

1 + 7— ll + y<br />

CO<br />

One can perform the multiplication in the denominator of (2.55) to obtain<br />

A>5 = ~A\s VSO /<br />

1 1<br />

+<br />

CO<br />

0>1 Q} 2 ) G>\1 *>2) (2.58)<br />

1<br />

s-A VSO<br />

l+y—<br />

o °>\


Analog Electronics /AmpliGers<br />

relates the approximate higher cut-off frequency fflb = (o H to the break frequencies CO, and 0)2. By<br />

using Equations (2.53) and (2.54) one obtains<br />

a> 0 s —— = - (2.60)<br />

r, + r 2 T 0<br />

It should be noted that Equation (2.60) underestimates the actual value of the higher cut-off<br />

frequency, which is the effect of dropping the right-side term in the denominator of Equation<br />

(2.56). The actual value of 0) H is higher than Oh and lower than minimum of CO, and co r It is then<br />

reasonable to write<br />

o) 0 H i,0)2} (261)<br />

The latter formula can be easily extended to three- and more-pole amplifiers.<br />

Using the numerical values assumed above for the elements of our circuit, we have A. = 265.2, C,<br />

= 2662 pF, C 2 = 10.04 pF, R, s 536 Q and R, = 5.3 kQ. They give z, = 1.426 us, r 2 = 52.83 ns, T 0 =<br />

1.479 us, which are respectively equivalent to co, = 701.36 krad/s, Q) 2 S 18.93 Mrad/s, and flfc =<br />

676.3 krad/s. Knowing that^ =


Analog Electronics /Amplifiers<br />

Zero phase characteristic of an amplifier results in an output waveform that is identical to the input<br />

(up to the scale factor). On the other hand, if the phase shift between the output and input signals<br />

of an amplifier is proportional to frequency, the output signal is a time-shifted version of the input,<br />

but their shapes are the same, so we do not say any distortions take place.<br />

Figure 2.26 Linear amplitude distortion: input signal (upper panel), output signal (lower panel).<br />

If the phase shift of an amplifier is not proportional to frequency, phase distortion occurs. As an<br />

example suppose the signal (2.63) is applied to uie inputs of three amplifiers A, B, C, which have a<br />

constant gain magnitude of 10. The amplifiers have different phase characteristics, as specified in<br />

Table 2.1<br />

TABLE 2.1 TRANSFER FUNCTION OF EXAMPLE AMPLIFIERS<br />

Amplifier GainatlkHz Gain at 3kHz<br />

A 10Z0° 10Z0°<br />

B 10Z-45 0 10Z-135 0<br />

C 10Z-45 0 10Z-45 0<br />

Apparently, amplifier A has a zero-phase response. The phase response of amplifier B is<br />

proportional to frequency with a proportionality constant of-45 degrees per kilohertz, and amplifier<br />

C has the phase response which is not proportional to frequency (or more precisely, it is nonzero<br />

constant, so the proportionality factor is zero). The plots of the output signals of the amplifiers are<br />

shown in Fig. 2.27. It is clear that the amplifier A does not change the shape of the signal.<br />

(However, it cannot be realized physically, as there is no delay between the input and the output -<br />

compare with Fig. 2.26.) Amplifier B also does not introduce any signal distortion, as the shapes of<br />

the input and output are the same. (Note the time delay between the signals.) Amplifier C that does<br />

not have its phase proportional to frequency produces severe distortion of the signal.<br />

We conclude that to avoid linear signal distortions, an amplifier should have constant gain<br />

magnitude and a phase response that is linear versus frequency for the range of frequencies<br />

contained in the input signal. For an audio amplifier it is required to have a constant gain for the<br />

frequency range from 20Hz to about 20kHz. However, since it turns out that the ear is not sensitive<br />

to phase distortion (at least for monophonic signals), we would not require the phase signals of an<br />

audio amplifier to be strictly proportional to frequency. Since the shape of the waveform ultimately<br />

determines the brightness of various points of an image, either amplitude or phase distortions of<br />

television signals would severely affect the image quality. Therefore we require the gain magnitude<br />

of a video amplifier to be constant and the phase response to be proportional to frequency in die<br />

whole range from dc to 6MHz.<br />

-38-


Analog Electronics /Amplifiers<br />

2.8 Pulse Response<br />

Often one needs to amplify a pulse signal such as the waveform v(l) shown in Fig. 2.28a. Pulses<br />

contain components that are spread over a wide range of frequencies; therefore amplification of<br />

pulses calls for a wideband amplifier. A typical amplifier output pulse is shown in Fig. 2.28a,<br />

denoted by v(7). The output waveform differs from the input in several important respects: the<br />

pulse is delayed, it displays overshoot and ringing, the leading and trailing edges are gradual rather<br />

than abrupt, and if the amplifier is ac-coupled, the top of the output pulse is tilted.<br />

The gradual rise of the leading edge of the amplifier response is quantified by giving the rise time /„<br />

which is the time interval between the point t 10 at which the amplifier achieves 10% of the eventual<br />

output amplitude K-and the point /j^ at which the output is 90% of the steady-state value. This is<br />

illustrated in Fig. 2.29.<br />

-39-


Analog Electronics/Amplifiers<br />

ho ho '<br />

Figure 2.29 Rising edge of a typical ac-coupied broadband amplifier output pulse<br />

(Note: no tilt is shown. When it is present, some judgement is necessary<br />

to estimate the finalamplitude Vj).<br />

The rounding of the leading edge can be attributed to the roll-off of gain in the high-frequency<br />

region. A rule-of-thumb relationship between the half-power bandwidth B and the rise time t r of a<br />

wideband amplifier is<br />

t r s ~Y (2-64)<br />

This relationship is not exact,for all types of amplifiers; it is accurate for firs' order (single pole)<br />

circuits. Analogously, the fall time U can be defined for the trailing edge of amplifier's pulse<br />

response. Equation (2.64) approximates the value of tr as well, at least for linear wideband<br />

amplifiers.<br />

Another aspect of the output pulse shown in Fig. 2.28 is overshoot and ringing, which are also<br />

related to the way the gain behaves in the high-frequency region. An amplifier that displays<br />

pronounced overshoot and ringing usually has a peak in its amplitude characteristic, as shown in<br />

Fig. 2.28b. The frequency of maximum gain approximately matches the ringing frequency. Because<br />

both rise time and overshoot are related to the high-frequency response, there is usually some tradeoff<br />

between these specifications. In a particular design, component values that reduce rise time<br />

often lead to more overshoot and tinging. Pulse amplifiers designed for fast rise time typically<br />

display about 10% overshoot because a higher amount of overshoot and associated ringing are<br />

usually undesirable.<br />

Figure 2.30 Pulse response that does not display ringing (a) and corresponding amplitude characteristic (b) of<br />

an amplifier<br />

Example 2.1 A television picture (25 pictures per second, 625 lines per picture) is produced by a<br />

TV camera that is able to reproduce about 765 distinguishable image elements within a 52^s active<br />

-40-


Analog Electtoaica/AxapHSets<br />

interval during each picture line. Estimate the bandwidth required for the video amplifier embedded<br />

within this camera.<br />

Solution. We estimate the bandwidth by the use of equation (2.22). Each picture element occupies<br />

die interval of 52)ls/765=68ns. Assume 80% of this interval is allowed to be the rise time: t r<br />

=54.4ns. Thus we obtain<br />

n 0.35 , A<br />

5 = -—— = 6.4 MHz<br />

54.4/w<br />

This is very close to die actual video bandwidth used in Europe.<br />

(2.65)<br />

Figure 2.31 Pulse responses (a) and amplitude characteristics (b) of ac-coupled amplifiers.<br />

(Tis the input pulse duration and r represents the shortest time constant of the coupling circuit)<br />

v(2): T« T, v(3): T=T, v(4): T» T.<br />

The tilt at the top of the output pulse, shown in Fig. 2.31a occurs if the amplifier is ac-coupled and<br />

originates from charging of coupling capacitors during the pulse. (After all, if the pulse lasted<br />

indefinitely, it would be the same as a new dc level at the input Since ac-coupled amplifiers do not<br />

pass me dc signal, output voltage would eventually return to zero.) Tilt (sag) is specified as a<br />

percentage of the initial pulse amplitude<br />

AF % = —-100% (2.66)<br />

where AV and V are defined in Fig. 2.32. As the duration of the input pulse increases (or die<br />

amplifier lower cut-off frequency is raised by changing the coupling circuits to have shorter time<br />

constants), output waveforms of consistently increased tilt result as shown in Fig. 2.31b.<br />

For small amount of tilt, the following approximate formula relates the percentage tilt to the lower<br />

cut-off (half-power) frequency^<br />

AV % =200^7 (2.67)<br />

where Tis the duration of the pulse.<br />

-41-


Analog Electronics /AmpliSera<br />

Figure 232 Pulse response that displays tilt<br />

Example 2.2 A cathode ray oscilloscope (CRO) is used to measure the pulse response of me<br />

amplifier shown in Figure 2.21. The equivalent impedance of the oscilloscope, including the cable<br />

connected between amplifies output and oscilloscope's input, is represented by a 1-MQ resistance<br />

in parallel with a 100-pF capacitance. This is illustrated in Figure 2.33. Find the output pulse rise<br />

time, as obtained from the measurement Compare with the value obtained for the circuit of Figure<br />

2.21.<br />

Figure 233 A transconductance amplifier with parallel capacitance and<br />

Oscilloscope + cable equivalent connected at the output<br />

Solution. We will use Equation (2.56). For the amplifier under consideration, the bandwidth equals<br />

to the higher cut-off frequency. To find the higher cut-off frequency for the circuits in Figure 2.21<br />

and 2.33, one can use Equation (2.35) with appropriate values for the equivalent resistance and<br />

capacitance to calculate the time constant In the case of Figure 2.21, we have<br />

— = —+ — (2.68)<br />

R o<br />

r o &L<br />

which for RL = 5.6 kQ and r 0 = 100 kfi gives Ro = 5.303 kfl. The equivalent capacitance is equal to<br />

Co = 10 pF. The time constant is the product RoG>, which is tti = 53.03 ns. This corresponds to the<br />

upper cut-off frequency of/k = 3 MHz and the rise time of £ = 0.12 ^ls. In the case of Figure 2.33,<br />

we have<br />

1 1 1 1<br />

• = — + • + -<br />

(2.69)<br />

Ro<br />

r o<br />

R L<br />

R CRO<br />

which for the assumed values gives Ro = 5.275 kQ. The equivalent capacitance is equal to (G> +<br />

CCRO) = HO pF. The time constant is the product RO(G)+CCRO), which is m = 580.2 ns. This<br />

corresponds to the upper cut-off frequency offii = 274 kHz and the rise time of U = 1.28 |J.s.<br />

One can see from the results of Example 2.2 that a substantial reduction of the apparent cut-off<br />

frequency can be experienced when measuring it by the oscilloscope, direcdy connecting die<br />

oscilloscope input to the amplifier output by a coaxial cable. Lower value of the cut-off frequency<br />

means that such an oscilloscope measurement setup introduces errors to the rise time and cannot<br />

be used to evaluate amplifier pulse response. The rise and fall times will apparency be much longer<br />

-42-


Analog Electronics /Amplifiers<br />

than their true values. To get rid of this undesired effect a device known as probe is used. The<br />

oscilloscope probe is an RC voltage attenuator as shown in Figure 2.34, where V^i denotes the<br />

voltage at die amplifier output and V2 is the voltage at the oscilloscope input Typically, a probe<br />

introduces 10 times attenuation of the input signal. If the probe capacitance Cp is properly adjusted,<br />

the attenuation is constant over a very wide frequency range. Such a probe is said to be<br />

compensated. For a compensated ptobe with 10-times attenuation, the input capacitance is 10<br />

times lower than the oscilloscope + cable capacitance. Thus the apparent reduction in the higher<br />

cut-off frequency is not as severe as it is with the CRO directly connected to the amplifier.<br />

Figure 234 Schematic diagram of the oscilloscope probe<br />

Example 2.3 Find the values of Rp and Cp of the oscilloscope probe in Figure 2.34 such that the<br />

attenuation V2/V\ is equal to 10 for all frequencies. Assume RCRO = 1 Mft and CCRO =100 pF.<br />

Solution. One can easily verify that the transfer function of the probe is described by<br />

A v =^ = P ^ n ( 2 - 70 )<br />

V l Rr, +R l + J*> R CRO C CRO ^<br />

P<br />

^R°<br />

1 + jaRpCp<br />

It will be a constant equal to<br />

Aw = *CRO (2J1)<br />

RCRO +<br />

R p<br />

for all frequencies if the following condition is satisfied<br />

RpCp = RcRO C CRO ( 2 - 72 )<br />

Putting Avo = 0.1 in Equation (2.71) one obtains.Kp = 9 MQ, which substituted into condition<br />

(2.72) gives Cp = 11.1 pF. It is suggested to the reader to find out that the input capacitance of the<br />

probe + oscilloscope arrangement is, under these circumstances, equal to Cm = 10 pF, whereas the<br />

parallel resistance Rin that loads the amplifier output is Rin = 10 MQ.<br />

Example 2.4 Find the upper cut-off frequency of the amplifier in Figure 2.33, as measured using a<br />

10-times attenuating compensated probe.<br />

Solution. We replace the oscilloscope equivalent RCROCCRO in Figure 2.33 by a parallel connection<br />

of Rin and Cm that represent the probe of Figure 2.34. Thus the equivalent output resistance of the<br />

amplifier is now equal to<br />

1 1 1 1<br />

— = — + — + — (2.73)<br />

R r 0 o<br />

R L Rin<br />

which for the assumed values gives Ro = 5.300 kfl. The equivalent capacitance is equal to (Co +<br />

Cm)= 20 pF. The time constant is the product Ro(G+Cm), which is m = 106 ns. This corresponds<br />

to the upper cut-off frequency of jk = 1.5 MHz and the rise time of £ = 0.23 ^s. The latter value is<br />

much closer to the true value of 0.12 |is than the rise time of 1.28 |4.s, as measured straight by the<br />

oscilloscope with no probe. This illustrates the advantage of using die probe for wideband amplifier<br />

characterization.<br />

-43-


Analog Electronics /Amplifiers<br />

2.9 Nonlinear Distortion<br />

Consider again an ideal resistive amplifier. Define its transfer characteristic as a ratio of the<br />

instantaneous output to instantaneous input For an ideal amplifier, the transfer characteristic can be<br />

plotted as a straight line on the output-input plane. For real amplifiers, diis characteristic remains<br />

linear over only a limited range of input and output voltages. For an amplifier operated from two<br />

power supplies the output voltage cannot exceed a specified positive limit and cannot decrease<br />

below a specified negative limit. Thus the output becomes saturated for large positive and large<br />

negative inputs. The resulting transfer characteristic is shown in Fig. 2.35, with the positive and the<br />

negative saturation levels denoted as L + and L_, respectively. Each of the two saturation limits is<br />

usually within 1 or 2 volts of the corresponding power supply. The output signal becomes distorted<br />

when me input exceeds the respective values 2, + / A y and L_ I A y . Obviously, in order to avoid<br />

distorting the output signal waveform, the input signal swing should be kept within the linear range<br />

of operation<br />

L_/A,Zv,ZLJA, (2.74)<br />

Fig. 2.35 shows the input waveforms (1) and (2) and the corresponding output waveforms. We note<br />

that the peaks of the larger waveform have been dipped off because of the saturation effect In<br />

some applications, like audio signal amplification, clipping is perceived as a rather severe distortion.<br />

Even small departures from linearity can be considered to be very serious in such cases. Assume the<br />

output-input relationship of a nonlinear amplifier can be written as<br />

v o (0 = M-(') + ^2[v,(')] 2 +4J[V,(01 3 +••• (2.75)<br />

where A u A& A 3 and so on are constants selected so diat the equation (2.75) matches the curvature<br />

of the nonlinear transfer characteristic. Obviously, for a linear amplifier, A 2 -A 3 =...= 0. Consider<br />

the case for which the input signal is a sinusoid given by<br />

v / (0 = ^ocos(fl> o r) (2.76)<br />

Now, we find an expression for the corresponding output signal. Substituting equation (2.76) into<br />

(2.75), applying trigonometric identities for [cos(flJ^]*, collecting terms and defining V 0 to be the<br />

sum of all of the constant terms, V, to be the sum of the terms with frequency 0)^, and so on, we<br />

find that<br />

v 0 (0 = V 0 Q + V ol cos(


Analog Electronics /Amplifiers<br />

The desired output is the V Q \ cos(co 0 t) term, which we call the fundamental component. The<br />

KJQ term represents a shift in the dc level (which does not appear at the load if it is ac-coupled). In<br />

addition, terms at multiples of the input frequency have resulted from the second and higher order<br />

nonlinear terms of the transfer characteristic (2.75). These terms are called harmonic distortion.<br />

The 2co 0 term is called second harmonic, the 3co 0 term is die third harmonic, and so on.<br />

Harmonic distortion is objectionable in a wideband amplifier because the harmonic can fall in die<br />

frequency range of the desired signal In an audio amplifier, harmonic distortion degrades die<br />

aesthetic qualities of the sound produced by the loudspeakers.<br />

The second harmonic distortion factor D 2 is defined as die ratio of the amplitude of the second<br />

harmonic to the amplitude of the fundamental.<br />

£) 2 =^2 (2.78)<br />

V ° l<br />

Similarly the third-order harmonic distortion factor, and so on, are defined as<br />

B-J-03 D 4 3 4 - (2.79)<br />

3<br />

v 0 { v ol<br />

The total harmonic distortion (THD) denoted by D is defined as<br />

D = VA 2 +^2+A 2 +- ( Z8 °)<br />

We can often find THD expressed as a percentage. A well-designed audio amplifier might have a<br />

THD specification of 0.01% (i.e. D- 0.0001) at rated output power.<br />

Notice diat THD specification of an amplifier depends of die amplitude of the output signal<br />

because the degree of nonlinearity of the transfer characteristic is amplitude-dependent. Certainly,<br />

any amplifier clips me output signal if the input becomes large enough. When clipping occurs, THD<br />

becomes large.<br />

Harmonic distortion is usually not a problem for bandpass amplifiers if the bandwidth is narrow<br />

enough so mat harmonics fall outside the frequency range of the desired signal. However, amplifier<br />

nonlinearity causes another type of distortion that can be very troublesome even for narrowband<br />

amplifiers. This is intermodulation and crossmodulation distortion that appear when die<br />

amplifier is excited by a sum of two or more sinusoidal components of close frequencies. Some of<br />

die intermodulation components generated due to amplifier nonlinearity fall into the original<br />

frequency band. The crossmodulation, which is the transfer of the amplitude of one input signal to<br />

a term wim a frequency of anodier input signal, is a very serious problem for radio receivers. These<br />

topics, however, fall outside the scope of this texdbook.<br />

Exercise 2.5 For the bipolar transistor amplifier shown below, write the PSpice code, run .AC<br />

analysis and plot amplitude and phase characteristics in the frequency range from 1Hz to 10MHz,<br />

100 points per decade. Identify the midband frequency range of this amplifier and find its halfpower<br />

bandwidth. Calculate its voltage, current and power gains at 1kHz. Explain me role of die<br />

capacitors C : and C 0 in the circuit<br />

-45-


Analog Electronics /Amplifiers<br />

I l><br />

R S CS | jg~""<br />

"••IT<br />

cc<br />

s<br />

Exercise 2.5- bipolar transistor common-amitter amplifier<br />

Vs 1 0 ac InV<br />

Rs 1 2 Ik<br />

Cs 2 3 luF<br />

RBI 7 3 330k<br />

RB2 3 0 100k<br />

Ql 5 3 4 Q2N2222A<br />

.LIB EVAL.LIB<br />

RE 4 0 10k<br />

CE<br />

RC<br />

Co<br />

RL<br />

4<br />

7<br />

5<br />

6<br />

470uF<br />

12k<br />

luF<br />

10k<br />

VCC 7 0 10V<br />

.ac dec 100 1Hz 10MEGHZ<br />

.end<br />

Exercise 2.6 For the bipolar transistor amplifier of Exercise 2.5, modify the PSpice code to run the<br />

transient and Fourier analyses for the time interval of 5ms, assuming that the input signal is a<br />

sinusoidal waveform of frequency 1 kHz and amplitude 1 mV. Repeat the simulation 10 times, each<br />

time for a different value of the input signal amplitude, covering the range from 2 mV to 20 mV.<br />

For each value of V s :<br />

(a) plot the waveforms of v(6) and -A*ik on a common diagram and compare their shapes,<br />

(b) find the corresponding value of THD, as recorded in the output file.<br />

(c) make the plot of the total harmonic distortion of the output signal as a function of input signal<br />

amplitude.<br />

2.10 Summary<br />

Linear amplifiers obey the- superposition principle. Their transfer characteristic, output voltage<br />

versus the input voltage, is a straight line with a slope equal to the voltage gain. Linear amplification<br />

can be obtained from a device having a nonlinear transfer characteristic by employing dc biasing<br />

and keeping the input signal amplitude small Depending on the signal to be amplified (voltage or<br />

current) and on the desired form of output signal (voltage or current), there are 4 basic amplifier<br />

types: voltage, current, transresistance and transconductance amplifiers. Amplifiers can be cascaded<br />

to increase the gain available, provided loading effects are taken into account Amplifiers increase<br />

-46-


Analog Electronics /Amplifiers<br />

the signal power and thus require dc power supplies for their operation, to convert the dc current<br />

energy from the power supply into the ac current energy of the signal amplified. Sinusoidal signals<br />

are invariant to linear operators; they are used to measure the frequency response of linear<br />

amplifiers. Amplifiers are classified according to the shape of their amplitude characteristic as<br />

lowpass, bandpass or highpass. The banpass amplifiers can be narrowband or wideband. Accoupled<br />

amplifiers are used to filter out dc component of the input signal and to break the dc<br />

current path between the signal source and the amplifier input, as well as between the amplifier<br />

output and the load. Ac-coupled amplifiers can never be classified as lowpass; they are either<br />

highpass or, more realistically, bandpass. In every physical amplifier there are always stray wiring or<br />

component capacitances in parallel with the signal path present At high frequencies they act as lowimpedance<br />

shunting elements that reduce the amplifier gain. Thus every physical amplifier has a<br />

limited bandwidth. The amplifier gain remains almost constant over the midfrequency band. It falls<br />

off at high frequencies where stray capacitances of components and devices no longer have high<br />

reactance. For ac amplifiers the gain falls off at low frequencies as well, because the coupling<br />

capacitors no longer have very low reactance. The amplifier bandwidth is the frequency range over<br />

which the gain remains within 3dB of the value at midband. The limits of the bandwidth are the<br />

frequencies^, and^. (For dc amplifier only^fr is meaningful). Variations of the amplitude and phase<br />

characteristics of an amplifier with frequency may cause distortion to the signal shape. Amplifiers<br />

do not distort signals if their gain is constant and phase is proportional to frequency for the range of<br />

frequencies contained in the input signal. If the frequency response of an amplifier is inadequate for<br />

a particular signal, there will be linear distortion — either amplitude distortion or phase distortion, or<br />

both. The shape of the unit step and pulse responses of the amplifier is related to its transfer<br />

function. Large bandwidth amplifiers produce pulses with a short rise time. Small values of the<br />

lower cutoff frequency give low percentage tilt. If the input signal amplitude is not sufficiently small,<br />

nonlinear distortion occurs due to inherent nonlinearities of amplifying devices. Nonlinear<br />

amplifiers do not obey the superposition principle. For sinusoidal input, sine waves of frequencies<br />

different to the input signal frequency appear in the output signal spectrum. Their amplitudes<br />

relative to the amplitude of the input sine-wave signal are the measure of nonlinear distortion<br />

introduced by an amplifier.<br />

-47-


Analog Electronics /Diodes<br />

3. Diode Circuits<br />

The simplest and most fundamental element of semiconductor circuits is the diode. Just like a<br />

resistor, it has two terminals; but unlike the resistor, it is a nonlinear element This chapter is<br />

concerned with the study of basic circuits that contain diodes. In order to understand the essence<br />

of diode function, we begin with a fictitious element, the ideal diode. We then introduce the real<br />

semiconductor junction diode, explain its terminal i-v characteristic and provide techniques for<br />

the analysis of diode circuits. The latter task involves the important subject of diode modeling,<br />

both for large-signal and small-signal applications.<br />

Of the many applications of diodes, their use in the design of rectifiers (which convert ac to dc) is<br />

the most common. Therefore, we shall study rectifier circuit in some detail and briefly look at a<br />

number of other diode applications.<br />

The semiconductor diode is a two-terminal device that incorporates a pn junction. The pn<br />

junction is the basis of many other solid-state devices, including the bipolar junction transistor.<br />

Thus an understanding of the/>» junction is essential to the study the material of this chapter.<br />

3.1 The Ideal Diode<br />

The ideal diode is a two-terminal device having the circuit symbol of Fig. 3.1(a) and the i-v<br />

characteristic shown in Fig. 3.1(b). One of its terminals is called the anode and the other is the<br />

cathode. The voltage v across the diode is referenced positive at the anode and negative at the<br />

cathode. The current is referenced positive when it flows from the anode to the cathode. The<br />

terminal characteristic of the ideal diode can be interpreted as follows: If a negative voltage is<br />

applied to the diode, no current flows and the diode behaves as an open circuit [Fig. 3.1(c)].<br />

Diodes operated in this mode are said to be reverse biased, or operated in the reverse direction.<br />

An ideal diode has a zero current when operated in the reverse direction and is said to be cut-off.<br />

(c)<br />

(d)<br />

Figure 3.1 The ideal diode: circuit symbol (a), i-v characteristic (b), equivalent circuit in the reverse<br />

direction (c), equivalent circuit in the forward direction (d).<br />

On the other hand, if a positive current is applied to the ideal diode, zero voltage-drop appears<br />

across the diode. In other words, the ideal diode behaves as a short circuit when biased in the<br />

forward direction [Fig. 3.1(d)]. It passes any current with zero voltage-drop. A forward-biased<br />

diode is called to be turned on or simply on.<br />

-48-


Analog Electronics /Diodes<br />

It is evident from Fig. 3.1(b) that the i-v characteristic is far away from a linear relationship. It<br />

contains two straight-line segments at 90 degrees one to another; thus it is highly nonlinear,<br />

although one may say it is piecewise linear. Other variations of piecewise-linear diode<br />

descriptions will be considered later in this chapter.<br />

'D<br />

V D 'D .. _ n l D =0<br />

v D =0<br />

VD<br />

(d)<br />

Figure 3.2 Rectifier circuit (a): equivalent circuit for v s >0 (b), equivalent circuit for v^0 (c),<br />

input waveform (d, upper panel),output waveform (d, lower panel).<br />

The fundamental application of the diode, one that makes use of its severely nonlinear i-v curve,<br />

is the rectifier circuit shown in Fig. 3.2a. The circuit consists of a series connection of a diode D<br />

and a resistor R. Let the input voltage be the sinusoid shown in Fig. 3.2d, and assume the diode<br />

to be ideal. During the positive half-cycles of the input sinusoid, the positive v s will cause current<br />

to flow through the diode in its forward direction. It follows that the voltage drop across the<br />

ideal diode will be zero. Thus the circuit will have the equivalent shown in Fig. 3.2(b), and the<br />

output )ltage v Q will be equal to the input voltage r s . On the other hand, during the negative<br />

half-cyciits of p f , the diode will not conduct. Thus the circuit will have the equivalent shown in<br />

Fig. 3.2c, and the output voltage will be zero. Note that while v s alternates in polarity and has a<br />

zero average value, v 0 is unidirectional and has a finite average value, or a non-zero dc<br />

component. Thus the circuit from Fig. 3.2a rectifiers the signal and hence is called a rectifier. It<br />

can be used to generate dc from ac. We will study more complex rectifier circuits later in this<br />

chapter.<br />

Exercise 3.1.For the circuit in Fig. 3.2a, sketch the transfer characteristic v Q versus v s .<br />

Exercise 3.2 For the circuit in Fig. 3.2a, sketch the waveform of the voltage across the diode.<br />

12V<br />

-49-


Analog Electronics /Diodes<br />

Example 3.1 Fig. 3.3 shows a circuit for charging a 12-V battery. If v s is a sinusoid with 24-V<br />

peak amplitude, find the fraction of each cycle during which the (ideal) diode is on. Also find the<br />

peak value of the diode current and the maximum voltage that appears across die diode and,<br />

finally, find the average current that charges the battery.<br />

Solution The diode conducts when s exceeds 12V, as shown in Fig. 3.3(b). The conduction<br />

angle is 20, where 6 is given by 24cos(6[)=12. Thus #=60° and the conduction angle is 120°, or<br />

one-third of a cycle. The peak value of the diode current occurs when the input voltage is<br />

maximum, and is given by 7 d =(24-12)/10 = 1.2A. The maximum reverse voltage across the diode<br />

occurs when v s is at negative peak and is equal to 24V+12V = 36V. The average current which<br />

1 *<br />

charges the battery is given as 1^ = —— J I d cos(g)d£ = 0.33 A, where £ is the integration<br />

variable.<br />

In<br />

1 IT *<br />

+5V<br />

00 (b)<br />

Figure 3.4 Diode logic gates: (a) OR gate; (b) AND gate<br />

Diodes together with resistors can be used to implement digital logic functions. Fig. 3.4 shows<br />

two diode logic gates. To see how these circuits function, consider a positive logic system in<br />

which voltage values close to 0 correspond to logic 0 (or low state) and voltage voltages close to<br />

+5V correspond to logic 1 (or high state). The circuit in Fig. 3.4a has three inputs, v A , v B , and v c .<br />

It is easy to see that diodes connected to +5V will conduct, thus clamping the output voltage v Y<br />

to a value equal to +5V. This positive output voltage will keep the diodes whose anodes are low<br />

(at zero voltage) cut off. Thus the output will be high if one or more inputs are high. The circuit<br />

in Fig. 3.4a therefore implements the logic OR function, which in Boolean notation is expressed<br />

as<br />

Y=A+B+C<br />

-50-


Analog Electronics<br />

/Diodes<br />

Similarly, the reader is encouraged to show that the circuit in Fig. 3.4b implements the logic<br />

AND function, that is<br />

Y = ABC.<br />

In analysis of circuits containing ideal diodes we may not know in advance which diodes are on<br />

and which are off. Thus we are forced to make a considered guess. Then we analyze the circuit to<br />

find the currents in the diodes assumed to be on and the voltage across the diodes assumed to be<br />

off. If the current i D through the diodes assumed to be on is positive and the voltage v D across the<br />

diodes assumed to be off is negative, our assumptions are correct and we have solved the circuit.<br />

Otherwise, me must make another assumption about the diodes and try again. After a little<br />

practice, our first guess is usually correct, at least for simple circuits.<br />

+I0V<br />

+I0V<br />

r-10V<br />

00 (b)<br />

Figure 3.5 Circuits for Example 3.2<br />

Example 3.2 Assuming the diodes to be ideal, find values of J and Kin the circuits of Fig. 3.5.<br />

Solution In these circuits it might not be obvious at first sight whether none, one or both diodes<br />

are conducting. In such a case we make a plausible assumption, proceed with the analysis and<br />

then check whether we end up with a consistent solution.<br />

For the circuit in Fig. 3.5a, we shall assume that both diodes are conducting. It follows that V B =0<br />

and V=0. The current through D2 can now be determined from<br />

I D2 = (10V-0V)/10k = 1mA.<br />

Writing the first Kirchhoff s law equation for node 8 we obtain<br />

1+lmA^ (0-(-l 0V))/5k = 2mA.<br />

Thus I=2mA-lmA = 1mA is larger then zero and thus diode Dl is conducting as originally<br />

assumed, and the final result is J=lmA, K=0.<br />

For the circuit in Fig. 3.5b, if we assume that both diodes are conducting, then V B =0 and V=0.<br />

The current through D2 is obtained from<br />

I D2 = (10V - 0V)/5k = 2mA<br />

The node equation at B is<br />

I+2mA = (OV - (-10V))/10k<br />

which yields I—1mA. The current is negative, so this result is not consistent with the assumption<br />

that diode Dl is on. We start again, assuming that Dl is off and D2 is on. The current through<br />

diode D2 is given by<br />

I D2 = (10V-(-10V))/(10k+5k)=1.33mA<br />

and the voltage at node B is<br />

V B = -10V +10k*1.33mA = 3.3V.<br />

Thus the voltage across Dl is negative (the anode is at 2ero and the cathode is at +3.3V), which<br />

means diode Dl is off as assumed, and the final result is 7=0 and K=3.3V.<br />

Exercise 3.3 Find the current I and voltage V for the circuits shown in Fig.3.6.<br />

Ans. (a) 2mA, 0V, (b) 0mA, 5V, (c) 0mA, -5V, (d) 2mA, 0V, (e) 3mA, 3V, (f) 4mA, IV.<br />

-51-


Analog Electronics /Diodes<br />

(e) (0<br />

Figure 3.6 Circuits for Exercise 3.3<br />

Exercise 3.4 Find the state of the ideal diodes in the circuits of Fig. 3.7<br />

lkQ<br />

4kO<br />

|6kfi<br />

5mA<br />

1 ©<br />

T3V<br />

10V<br />

Figure 3.7 Circuits for Exercise 3.4<br />

Exercise 3.5 Fig. 3.8 shows a circuit for an ac voltmeter. It utilizes a moving-coil meter that<br />

gives a full-scale reading when the average current flowing through it is 1mA. Hie moving-coil<br />

meter has a 50 Q resistance. Find the value R that results in the meter indicating full-scale reading<br />

when the input sine-wave voltage r s is 20V peak-to-peak. Ans. R=3.133 kfi.<br />

Moving-coil<br />

'oJ meter<br />

Figure 3.8 Circuit for Exercise 3.5<br />

3.2 Terminal Characteristics of Semiconductor Diodes<br />

In this section we study the characteristics of real diodes - specifically semiconductor junction<br />

diodes made of silicon. Fig. 3.9 shows an example of the i-v characteristic of a silicon junction<br />

diode. As indicated, the characteristic curve consists of three distinct regions:<br />

1. The forward-bias region determined by v>0.<br />

o—W-<br />

-52-


Analog Electronics / Diodes<br />

2. The reverse-bias region, determined by v


Analog Electronics /Diodes<br />

v = nV T \n{—) (3.4)<br />

The exponential relationship of the current / to the voltage v holds over many decades of current<br />

(a span of as many as seven decades - that is a factor of 1CK 7 - can be found). This is quite a<br />

remarkable property of junction diodes (and bipolar transistors as well), one that has been<br />

exploited in many interesting applications. The diode i-v characteristic is then most conveniently<br />

plotted on a semilog paper. One can show that using the vertical, linear axis for v and the<br />

horizontal, log axis for i, one obtains from (3.3) a straight line with a slope of 2.3«K T per decade<br />

of current.<br />

Since both I s and V T in (3.3) are functions of temperature, the forward i-v characteristic varies<br />

with temperature as illustrated in Fig. 3.10. At a constant diode current, the voltage drop across<br />

the silicon diode decreases by approximately 2mV for every kelvin increase in temperature. This<br />

property has been exploited in the design of electronic thermometers.<br />

A glance at the i-v characteristic in the forward region (Fig. 3.9) reveals that the current / is<br />

negligibly small for v smaller than about 0.5V. This value is usually referred to as the knee<br />

voltage. It should be emphasized, however, that this apparent threshold in the diode<br />

characteristic is simply a consequence of the exponential relationship. Another consequence of<br />

this relationship is the rapid increase of /'. Thus for a "fully conducting" diode; the voltage drop<br />

across it lies in a narrow range, approximately from 0.6 to 0.8V for silicon diodes. This gives rise<br />

to a simple model of a diode where it is assumed that a conducting diode is replaced by a 0.7-V<br />

battery. This constant-voltage-drop model is discussed later on in this chapter. Diodes with<br />

different ratings (i.e. different areas and different 1$) will exhibit the 0.7-V drop at different<br />

currents. For instance, a low-power switching diode may be considered to have 0.7V drop at<br />

/'=lmA, while a higher power diode may have a 0.7V drop at /-1A.<br />

-54-


Analog Electronics /Diodes<br />

The reverse-bias region. The reverse-bias region is entered when the diode voltage v is made<br />

negative. Equation (3.1) predicts that if v is negative and its absolute value is a few times larger<br />

than V T (25mV), the exponential term becomes negligibly small compared to unity and the diode<br />

current becomes<br />

i=-Is (3.5)<br />

that is, the current in the reverse direction is constant and equal to I s -.'m magnitude. This is the<br />

reason behind the term saturation current: the current saturates when-1 v\>Vj, v


Analog Electronics /Diodes<br />

The second mechanism causing breakdown is called the Zener effect. It occurs in abrupt<br />

junctions having high doping levels. For such junctions, the depletion region is very narrow<br />

because die charge density of ionized dopant atoms is high and a large amount of charge can be<br />

stored in a very thin layer. As reverse bias is applied, the field in the depletion region increases in<br />

intensity. When the field strength is of the order of 1 V divided by the crystal lattice spacing, it is<br />

possible for covalent bonds to be broken by the field. In other words, the forces are so strong<br />

that electrons are pulled loose from their bonds.<br />

The energy gap between the top of the valence band and the bottom of the conduction band<br />

becomes slightly smaller with increased temperature. Hence the force required to break the<br />

covalent bonds is slighdy smaller at higher temperatures. As a result, if the Zener effect is<br />

responsible for breakdown, the breakdown voltage tends to become smaller with increased<br />

temperature. This is opposite to the case for avalanche breakdown.<br />

As a rule, the Zener effect is responsible if the breakdown voltage magnitude is less man 6 V, and<br />

the avalanche effect is responsible if the breakdown voltage is greater than 6 V. For diodes<br />

having breakdown voltages of approximately 6 V, a mixture of both mechanisms can occur. It is<br />

possible to obtain diodes having breakdown voltages of about 6 V with very small temperature<br />

coefficients, because the temperature effects of the two breakdown types offset one another.<br />

Dynamic impedance also tends to be a function of breakdown voltage, reaching a minimum for<br />

breakdown voltages of about 6 V. Thus, as circuit designers, we tend to select 6-V diodes.<br />

As can be seen from Fig. 3.9, in the breakdown region the reverse current increases rapidly, and<br />

the associated increase in the voltage drop is very small. Diode breakdown is normally not<br />

destructive, provided that the power dissipated in the diode is limited by external circuitry to a<br />

"safe" level. This safe value is normally specified on the device data sheets. It therefore is<br />

necessary to limit the reverse current in the breakdown region such that the product of the dc<br />

current and the dc voltage is lower than the permissible power dissipation.<br />

The fact that the diode i-v characteristic in breakdown is almost a vertical line enables it to be<br />

used in voltage regulation (stabilization). This subject will be studied in Section 3.6.<br />

3.3 Analysis of Diode Circuits<br />

In this section, we shall study methods for the analysis of diode circuits. We shall concentrate on<br />

circuits in which the diodes are operating in the forward-bias region. Operation in the breakdown<br />

region is considered in Section 3.5.<br />

Consider the circuit shown in Fig. 3.11 consisting of a dc source V DDJ a resistor R and a diode.<br />

We wish to analyze this circuit to determine the diode current I D and the diode voltage V D .<br />

R ID<br />

Figure 3.11 A simple diode circuit<br />

-56-


Analog Electronics /Diodes<br />

The diode is obviously biased in the forward direction. Assuming that V DD is greater than 0.1V or<br />

so, the diode current will be much greater than I s and we can represent the diode i-v characteristic<br />

by the exponential relationship<br />

/ D = / S exp(-^-) (3.6)<br />

The other equation that governs the circuit operation is obtained by writing a Kirchhoff loop<br />

equation, resulting in<br />

I D =(V DD -V D )/R (3.7)<br />

Assuming that the diode parameters I s and n are known, Eqs. (3.6) and (3.7) are two equations in<br />

the two unknown quantities I D and Vp. Two alternative ways for obtaining the solution are<br />

graphical analysis and iterative analysis.<br />

Graphical analysis. Graphical analysis is performed by plotting the relationship (3.6) and (3.7)<br />

on the i-v plane. The solution can then be obtained as the coordinates of the point of intersection<br />

of the two graphs. A sketch of the graphical construction is shown in Fig. 3.11; the curve<br />

represents the exponential diode equation (3.6) and the straight line represents (3.7). Such a<br />

straight line is known as the load line, a name that will become more meaningful in later<br />

chapters. The load line intersects the diode curve at point Q, which represents the operating<br />

point of the circuit. Its coordinates give the values of I D and V D .<br />

Graphical analysis aids in the visualization of circuit operation. However, the effort involved in<br />

performing such an analysis, particularly for complex circuits, is too great to be justified in<br />

practice.<br />

V DD /R<br />

0-point<br />

o v D v DD<br />

Figure 3.12 Graphical analysis of the circuit in Fig. 3.11<br />

Iterative analysis. Equations (3.6) and (3.7) can be solved using a simple iterative procedure, as<br />

the one illustrated in the following example.<br />

Example 3.3 Determine the current I D and the diode voltage V D for the circuit in Fig. 3.11 with<br />

K DD =5V and R=lk. Assume that^lO 10 , «=1.7.<br />

Solution Taking a look at Figure 3.11 one may arrive at conclusion that the diode is biased into<br />

the forward direction. Thus Equation (3.6) describes the diode current Taking a natural<br />

logarithm of both sides of (3.6) one obtains<br />

V D =»V T ki(I D /Q.<br />

Now, substituting (3.7) for I D in the above formula gives<br />

V D =»V T ]n((V DD -V D )/(RI s ))<br />

The last equation can be rewritten for iterative calculations, such that the diode voltage in<br />

iteration (k+1) is related to its value in the previous iteration (k) as follows<br />

V D W = »V T ]n((V DD -V D


Analog Electronics<br />

/Diodes<br />

Assume the initial guess tor the diode voltage is V D<br />

—0.5V. Its values and the corresponding<br />

values of the current in successive iterations are<br />

k vj> J «<br />

0 0.5V 0.013mA<br />

1 0.748942V 4.500mA<br />

2 0.746524V 4.251mA<br />

3 0.746524V 4.251mA<br />

4 ...<br />

Thus the third iteration yields J D =4.251mA and K D =0.74624V. Since these values are not much<br />

different from the values obtained after the second iteration, no further iterations are necessary.<br />

Figure 3.13 The effect of diode series resistance<br />

The iterative analysis procedure utilized in the example above is simple and yields accurate results<br />

after two or three iterations. Nevertheless, there are situations in which the effort and time<br />

required to perform the iterative calculations are still greater than can be justified. Specifically,<br />

when one is doing a pencil-and-paper design of a relatively complex circuit, rapid circuit analysis<br />

is a necessity. Through quick analysis, the designer is able to evaluate various possibilities before<br />

deciding on a suitable circuit. To speed up the analysis process, one must be content with less<br />

precise results. This, however, is seldom a problem, as the more accurate analysis can be<br />

postponed until a final or almost final design is obtained. Accurate analysis of the final or almost<br />

final design can be performed with the aid of a computer-analysis program such as SPICE. The<br />

results of such an analysis can then be used to further refine or "fine-tune" the design.<br />

Diode models. At high current levels, the ohmic resistance of the semiconductor forming the<br />

junction becomes significant. Addition of a series resistance R t to the diode modeled by the<br />

Shockley equation (3.1) can account for this. The modified version of (3.4) becomes<br />

-58-


Analog Electronics /Diodes<br />

v-«F I .In(7-) + « J<br />

(3.8)<br />

Typical low-current diodes have Revalues ranging from 10 to 100Q. The resistance R, decreases<br />

with the diode cross-sectional area. Thus for high-current diodes it is of the order of 0.1Q. The<br />

effect of die diode series resistance on the i-v characteristic is illustrated in Fig. 3.13. For a fixed<br />

value of current, the voltage drop across the diode external terminals increases with the series<br />

resistance value.<br />

Slope=l/r£)<br />

(c)<br />

Figure 3.14 Approximating the diode forward characteristic with two straight lines (a), (b);<br />

equivalent circuit representation of the piecewise-linear model (c).<br />

Although the exponential i-v characteristic plus a series resistance is an accurate model in the<br />

forward region, its nonlinear nature complicates the analysis of diode circuits. The analysis can be<br />

greatly simplified if we can find piecewise-linear relationship to describe the diode terminal<br />

characteristics. An attempt in this direction is illustrated in Fig. 3.14, where the exponential curve<br />

is approximated by two straight-line segments, line A with zero slope and line B with a slope of<br />

//r D . It can be seen that for this particular diode, over the current range from 0 to 10mA the<br />

voltages predicted by the straight-line model differ from those predicted by the exponential<br />

model by less than 50 mV. Of course, the choice of two lines is not unique; one can obtain a<br />

closer approximation by restricting the current range over which the approximation is required.<br />

The straight-lines (or piecewise-linear) model of Fig. 3.14 can be described by<br />

»D = °><br />

V D ^ VDo<br />

h =<br />

Do<br />

v D >V Do (3.9)<br />

where V D , is the intercept of line B on the voltage axis and r D is the inverse of the slope of line B.<br />

For the particular example shown, V D , = 0.65 V and r D = 20 Q.<br />

-59-


Analog Electronics /Diodes<br />

The equivalent circuit shown in Fig. 3.14c can represent the piecewise-linear model described by<br />

(3.8). Note that an ideal diode is included in this model to constrain i D to flow in the forward<br />

direction only. This model is also known as the "battery-plus-resistance" model.<br />

Example 3.4 Repeat the problem in Example 3.3 utilizing the piecewise-linear model whose<br />

parameters are V Dt — 0.65 V, r D = 20 Q.<br />

Solution Replacing the diode in Fig. 3.11 with the equivalent circuit model of Fig. 3.14c results in<br />

the circuit in Fig. 3.15, from which we can write the current I D<br />

ID = {VDD - VJ/(R + r D ) =4.26mA<br />

The diode voltage can now be computed<br />

VD = VD. + roI D = 0.7353V<br />

Note that the values obtained using the simplified model are not very much different from the<br />

accurate values obtained in Example 3.3.<br />

Figure 3.15 The circuit of Fig. 3.11 with the diode replaced by its<br />

piecewise-linear model of Fig. 3.14c.<br />

An even simpler model of the diode forward characteristic can be obtained if we use a vertical<br />

straight line to approximate the fast-rising part of the exponential curve, as shown in Fig. 3.16a.<br />

The resulting model simply says that a forward-conducting diode exhibits a constant voltage drop<br />

V D . The value of V D for silicon diodes is usually taken to be 0.7V. The equivalent circuit shown<br />

in Fig. 3.16c can represent the constant-voltage-drop model. This model is the one most<br />

frequently employed in initial phases of analysis and design. Finally, note that if we employ the<br />

constant-voltage-drop model to solve the problem in Examples 3.3 and 3.4 we obtain<br />

IQ = (V DD -V D )IRI D = 4.3 mA which is not too differ^ it from the values obtained with the<br />

more elaborate models.<br />

V, D<br />

(b)<br />

-60-


Analog Electronics /Diodes<br />

Ideal<br />

V D<br />

(c)<br />

Figure 3.16 Development of the constant-voltage-drop model of the diode forward characteristic with<br />

two straight lines (a), (b); equivalent circuit representation of the model (c).<br />

In applications that involve voltages much greater than the diode voltage drop (0.6 - 0.8V), we<br />

may neglect the diode voltage drop altogether while calculating the diode current The result is<br />

die ideal diode model, which we studied in Section 3.1.<br />

The question of which diode model to use in a particular application is one a circuit designer<br />

faces repeatedly, not just with diodes but with every circuit element. One's ability to select<br />

appropriate device models improves widi practice and experience.<br />

3.4 The Diode Small-Signal Model at Low Frequencies<br />

We will encounter many examples of electronic circuits in which dc supply voltages are used to<br />

bias a nonlinear device at an operating point and a small ac signal is injected into the circuit. We<br />

often split analysis of such circuits into two parts. First, we analyze the dc circuit to find the<br />

operating point. In this analysis of bias conditions, we must deal with die nonlinear aspects of the<br />

device. In the second part of the analysis, we consider the small ac signal. Since virtually any<br />

nonlinear characteristic is approximately linear (straight) if we consider a sufficiendy small<br />

portion, we can find a linear small-signal equivalent circuit for the nonlinear device to use in<br />

die ac analysis.<br />

The small-signal linear equivalent circuit is an important analysis approach that applies to many<br />

types of electronic circuits. In diis section we demonstrate the principles with a simple diode<br />

circuit. In the next chapter we use similar techniques for transistor amplifier circuits.<br />

Now we will show that in the case of a diode, the small-signal equivalent circuit consists of a<br />

resistance. Consider a conceptual circuit in Fig. 3.17a and the corresponding graphical<br />

representation in Fig. 3.17b. A dc voltage Kp, represented by a battery, is applied to the diode;<br />

and a time-varying signal iy(/), assumed (arbitrarily) to have a triangular waveform, is added to<br />

(superimposed on) the dc voltage V D . In the absence of the signal ty(/) die diode voltage is equal<br />

to Vp, and correspondingly the diode will conduct the current I D given by the Shockley equation<br />

(3.1), which for V D »V T , (i.e. for sufficiently large forward current) simplifies to<br />

/ D = / S exp(-^T) (3-10)<br />

When die signal ta(/) is applied, the total instantaneous diode voltage v D (t) will be given by<br />

v D (t) = V D +v d (t) (3.11)<br />

Correspondingly, the total instantaneous diode current i D (t) will be<br />

K ' +V < ( 'fl (3,2,<br />

which can be rewritten<br />

-61-


Analog Electronics /Diodes<br />

Figure 3.17 Development of a diode small-signal model<br />

Using (3.10) we obtain<br />

/ D (0 = / D exp(^J<br />

(3.14)<br />

Now if the magnitude of the signal vt(t) is kept sufficiently small such that<br />

v,(0<br />

« 1<br />

(3.15)<br />

nV T<br />

then we may expand the exponential of (3.14) into the Taylor series and truncate the series after<br />

the first two terms to obtain approximate expression<br />

V ' ( '^ (3.16)<br />

i D (0 = I D ! + • nVT J<br />

This is the small-signal approximation. It is valid for signals whose amplitudes are smaller than<br />

about lOmV.<br />

From (3.16) we have<br />

i D (0 = I D +^-v d (t) (3.17)<br />

Thus superimposed on the dc current I D we have a signal component directly proportional to the<br />

signal voltage m(/). That is<br />

where<br />


Analog Electronics / Diodes<br />

r


Analog Electronics /Diodes<br />

Notice that the ac signal to be attenuated is connected to the circuit by a coupling capacitor.<br />

The output voltage is connected to the load R^ by a second coupling capacitor. However, the<br />

coupling capacitors are open circuits for dc. Thus die j2-point of die diode is unaffected by the<br />

signal source or the load. Furthermore, the coupling capacitors prevent (sometimes-undesirable)<br />

dc currents from flowing in the source or the load.<br />

Figure 3.19 Variable attenuator using a diode as a controlled resistance<br />

Because of the coupling capacitors, we only need to consider V c , R^ and the diode to perform<br />

the dc analysis to find the =


Analog Electronics / Diodes<br />

Figure 3.21 Small signal equivalent circuit for Figure 3.19<br />

Exercise 3.6 Suppose that the circuit of Fig. 3.19 has R = 100 Q, R^ - 2 kQ, and R L = 2 kQ.<br />

The diode has n - land is at a temperature of 300K. For purposes of j2-point analysis assume a<br />

constant diode voltage of 0.6 V. Find the j2-point value of the diode current and A vs for (a) V c -<br />

1.6 V and (b) V c = 10.6 V. Ans. (a) 1 D = 0.5 mA,A v = 0.331, (b) 7 D = 5*1^,^ = 0.0492.<br />

An application for voltage-controlled attenuator occurs in tape recorders, as an example. A<br />

problem frequently encountered in recording a conversation is that some persons speak quietly,<br />

whereas others speak loudly. Furthermore, some may be far from the microphone, whereas<br />

others are close. If an amplifier with fixed gain is used between the microphone and the tape<br />

head, either the weak signals are small compared the noise level or the strong signals drive the<br />

recording nonlinear so that severe distortion occurs.<br />

Figure 3.22 The voltage-controlled attenuator is useful in maintaining a<br />

suitable signal amplitude at the recording head<br />

A solution is to use a voltage-controlled attenuator placed between the microphone and a highgain<br />

amplifier in a system like that shown in Fig. 3.22. When the signal being recorded is weak,<br />

the control voltage is small and very little attenuation occurs. On the other hand, when the signal<br />

is stron^, the control voltage is large, so the signal is attenuated, preventing distortion. Rectifying<br />

the output of the amplifier generates the control voltage. The rectified signal is filtered by a longtime-constant<br />

RC filter so that the attenuation responds to the long-term average signal<br />

amplitude rather than adjusting too rapidly. With proper design, this system can provide an<br />

acceptable signal at the recording heads for a wide range of input signal amplitudes. Eventually,<br />

we will discuss all of the circuits required in this system.<br />

-65-


Analog Electronics /Diode Circuits<br />

3.5 Rectifier Circuits<br />

Now that we have introduced the diode and some methods for analysis of diode circuits, we<br />

consider some practical circuits. First, we consider several types of rectifiers that convert ac<br />

power into dc power. These rectifiers form the basis for electronic power supplies and batterycharging<br />

circuits. Other applications for rectifiers are in signal processing, such as demodulation<br />

of a radio signal, or measuring the average amplitude of a signal from a microphone, as<br />

considered in the previous section. Another application is precision conversion of an ac signal to<br />

the dc in an electronic voltmeter.<br />

Half-wave rectifier circuits. A half-wave rectifier with a sinusoidal source and resistive load is<br />

shown in Fig. 3.23a. (The same circuit has already been analyzed in Section 3.1, for the ideal<br />

diode case.). When the source voltage is positive, the diode is in the forward-bias region. If an<br />

ideal diode is assumed, the source voltage appears across the load. For a typical real diode, the<br />

output voltage is less than the source voltage by an amount equal to the drop across the diode,<br />

which is about 0.7V for silicon diodes at "room temperature". This can be seen in Fig. 3.23b.<br />

(a)<br />

(b)<br />

Figure 3.23 Half-wave rectifier with resistive load (a); input and load voltages versus time (b)<br />

If the source voltage is negative, the diode is reverse b;.. :d and no current flows through the<br />

load. Even for typical real diodes only a very small reverse current flows. Thus only the positive<br />

half-cycles of the source voltage appear across the load. One can use a half-wave rectifier to<br />

charge a battery, as shown in Fig. 3.3a. Current flows in that circuit whenever the instantaneous<br />

ac source voltage is higher than the battery voltage. As shown in the figure, it is often necessary<br />

to add resistance in series with the diode to limit the magnitude of the current. When the ac<br />

source voltage is less than'the battery voltage, the current is zero. Thus the current flows only in<br />

die direction that charges the battery.<br />

Half-wave rectifier with smoothing capacitor. Often, we want to convert an ac voltage into a<br />

nearly constant dc voltage to be used as a power supply for our designs. One approach to<br />

smoothing the rectifier output is to place a large capacitor across the output terminals of the<br />

rectifier. The circuit and waveforms of current and voltage are shown in Fig. 3.24. When die ac<br />

source reaches a positive peak, the capacitor is charged to the peak voltage (assuming an ideal<br />

diode). Then when the voltage drops below the voltage stored on the capacitor, the diode is<br />

reverse biased and no current flows through the diode. The capacitor continues to supply current<br />

to the load, slowly discharging until the next positive peak of the ac input. As shown in the<br />

figure, current flows through the diode in pulses that recharge the capacitor.<br />

-66-


Analog Electronics /Diode Circuits<br />

« (b)<br />

Figure 3.24 Half-wave rectifier with smoothing capacitor (a); voltage and current waveforms (b)<br />

Because of the charge and discharge cycle, the load voltage contains a small ac component called<br />

ripple. Usually, it is desirable to minimize ripple, so we choose the largest capacitance value that<br />

is practical. In this case, the capacitor discharges for nearly the entire cycle, and the charge<br />

removed from the capacitor during one discharge cycle is<br />

QsJJ (3.26)<br />

where I L is the average load current and T is the period of the ac voltage. Since the charge<br />

removed from the capacitor is the product of the change in voltage and the capacitance, we can<br />

also write<br />

Q=V r C (3.27)<br />

where V r is the peak-to-peak ripple voltage and C is the capacitance. Equating the right-hand<br />

sides of (3.27) and (3.26) allows one to solve for C<br />

/ T<br />

c=-£- (3.28)<br />

In practice, Equation (3.28) is approximate because the load current varies in time and because<br />

the cap; itor does not discharge for a complete cycle. However, it gives a good-starting value for<br />

the capacitance required in the design of power-supply circuits. We will return to the subject of<br />

power-supply design after we have introduced transistor circuits and operational amplifier<br />

feedback circuits.<br />

The average voltage supplied to the load if a smoothing capacitor is used is approximately<br />

midway between the minimum and maximum voltages. Thus, referring to Fig. 3.24, the average<br />

load voltage is<br />

V.<br />

v L =v m -v D - (3.29)<br />

An important aspect of rectifier circuits is the peak inverse voltage (PIV) across the diodes. Of<br />

course, the breakdown specification of the diodes should be greater in magnitude than PIV. For<br />

example, in the half-wave circuit with a resistive load, shown in Fig. 3.23, the PIV is V m .<br />

Addition of a smoothing capacitor in parallel with the load increases the PIV to (approximately)<br />

2V m . Referring to Fig. 3.24, for the negative peak of the ac input, we see that the reverse bias of<br />

the diode is the sum of the source voltage and the voltage stored on the capacitor.<br />

Full-wave rectifier circuits. Several full-wave rectifiers are in common use. One approach<br />

uses a center-tapped transformer and two diodes as shown in Fig. 3.25a. This circuit consists of<br />

-67-


Analog Electronics/Diode Circuits<br />

two half-wave rectifiers widi out-of-phase source voltages and a common load. The diodes<br />

conduct on alternate half-cycles.<br />

Figure 3.25 Center-tapped-transformer full-wave rectifier<br />

Besides providing the out-of-phase ac voltages, the transformer also allows adjustment of V m by<br />

selection of turns ratio. This is an important function because the ac voltage available is often not<br />

suitable for direct rectification - usually either higher or lower dc voltage is required.<br />

A second type of full-wave rectifier uses the diode bridge shown in Fig. 3.26a. When the ac<br />

voltage is positive at the top of the secondary winding, current flows through diode A, then<br />

through the load and returns through diode B as shown in the figure. For the opposite polarity,<br />

i.e. during the alternate half-cycle, current flows through diodes C and D. Notice that in either<br />

case the current flows in the same direction through the load. For real diodes, the peak voltage<br />

across the load is lower in the case of the diode bridge rectifier circuit compared to the centertapped<br />

transformer rectifier circuit. This is because two diodes are connected in series with the<br />

load in the diode bridge circuit and, consequently, twice the voltage drop across a forward-biased<br />

diode is subtracted from the ac source voltage to obtain the load voltage. A single voltage drop<br />

across the diode is subtracted in the case of the circuit in Fig. 3.25a. This property is illustrated in<br />

Fig. 3.26b and Fig. 3.25b.<br />

If one side of die load is connected to die ground as shown in Fig. 3.26a, neither of the ac source<br />

terminals of the diode bridge can be connected to the ground. The transformer with floating<br />

secondary winding helps to realize this condition. (If both the ac source and the load had a<br />

common ground connection and no transformer was used, part of die circuit would be shorted.)<br />

If one wishes to smooth the voltage across the load, a capacitor can be placed in parallel with the<br />

load similar to the half-wave circuit discussed earlier. In the full-wave circuits, the capacitor<br />

-68-


Analog Electronics /Diode Circuits<br />

discharges for only a half-cycle before being recharged. Thus the capacitance required<br />

maintaining a given level of ripple is only half as much in the full-wave circuit as for the halfwave<br />

circuit. Therefore, Equation (3.28) can be modified to obtain<br />

C = - L -<br />

IV.<br />

for the full-wave rectifier with a capacitor filter.<br />

(3.30)<br />

-5.9U +<br />

*•<br />

n UUMIU)<br />

•. UC2)<br />

(b)<br />

Fig. 3.26 Diode-bridge full-wave rectifier<br />

Power supply circuit. A diode rectifier forms an essential building block of the dc power<br />

supplies required to feed electronic equipment. A block diagram of such a power supply is shown<br />

in Fig. 3.27. As indicated, the power supply is fed from the 220-V (tms) 50-Hz ac line, and it<br />

delivers a dc voltage V L (usually in the range of 5 to 20V) to an electronic circuit represented by<br />

the load block. The dc voltage V L is required to be as constant as possible in spite of variations<br />

in the ac line voltage and in the current drawn by the load. Normally, the ripple at the output of a<br />

rectifier circuit, with a smoothing capacitor of practical value, are to big and do not fulfill the<br />

requirements of most electronic equipment Further means are necessary to stabilize (regulate)<br />

the voltage across the load.<br />

-69-


Analog Electronics /Diode Circuits<br />

The first block in the power supply is the power supply. It consists of two separate coils wound<br />

around an iron core that magnetically couples the two windings. The primary winding, having<br />

N 7 turns, is connected to the 220-V ac supply. The secondary winding, having N 2 turns, is<br />

connected to the rectifier circuit We denote the turns ratio by n=N 1 /N 2 . Thus an ac voltage v s of<br />

220/ ft V(rms) develops between the two terminals of the secondary winding. By selecting the<br />

appropriate turns ratio n for the transformer, the designer can step the line voltage down to the<br />

value required to yield the particular dc voltage output of the supply. For instance, a secondary<br />

voltage of 8-V rms is usually required for a dc output of 5V. This can be achieved with the<br />

«=28:1 turns ratio.<br />

It should be noted that to simplify the discussion, zero-resistance transformer windings ate<br />

assumed for the rectifier circuits analyzed in this Section. Thus the results obtained are<br />

approximate only; however, they can still be sufficiently accurate for low-power applications.<br />

In addition to providing the appropriate sinusoidal amplitude for the dc power supply, the power<br />

transformer provides electrical isolation between the electronic equipment and die power line<br />

circuit. The isolation minimizes die risk of electric shock to the equipment user.<br />

The diode rectifier converts the input sinusoid v s to a unipolar output that can have a pulsating<br />

waveform indicated in Fig. 3.27. Although this waveform has a nonzero average, or a dc<br />

component, its pulsating nature makes it unsuitable as a dc source of electronic equipment, hence<br />

die need for a filter that suppresses high-frequency components of the rectifier's signal In its<br />

simplest form, the filter is a smoothing capacitor, as discussed above. The filter block in Fig. 3.27<br />

significantly reduces the variations of the rectifier output.<br />

The filtered rectifier output, though much more constant than without the filter, still contains a<br />

time-dependent component, known as ripple. To reduce the ripple and to stabilize the magnitude<br />

of the dc output voltage of the supply against variations caused by changes of in load current, a<br />

voltage regulator is employed. Such a regulator can be implemented using a Zener diode, which<br />

will be discussed, in the following section. Alternatively, an integrated circuit (IC) regulator can<br />

be used.<br />

Exercise 3.7 A power-supply circuit is needed to deliver 0.1 A and 15V (average) to the load. The<br />

ac source available is 220V rms with a frequency of 50Hz. Assume that the full-wave circuit of.<br />

Fig. 3.26 is to be used with a smoothing capacitor in parallel with the load. The peak-to-peak<br />

ripple voltage is to be 0.4V. Allow 0.7V for forward diode drop. Find the turns ratio n needed<br />

and the approximate value of die smoothing capacitor. Verify your results using SPICE. Ans.<br />

»=18.74, C=2500nF.<br />

Exercise 3.8 Repeat Exercise 3.7 using the circuit of Fig. 3.25 with a smoothing capacitor in<br />

parallel with the load resistance. (Define the turns ratio as the ratio of primary turns to the<br />

secondary, turns between the center tap and one end.) Verify your results using SPICE. Ans.<br />

»=19.58, C=2500nF.<br />

3.6 Zenor-Diode Voltage Regulator Circuits<br />

The circuit shown in Fig. 3.28 is used to provide a nearly constant output voltage from a variable<br />

source. (For its proper operation, it is necessary for the minimum value of the variable source<br />

voltage to be somewhat larger than the desired output voltage.) A Zener diode having a<br />

breakdown voltage equal to the desired output voltage is used. The resistor R limits die diode<br />

-70-


Analog Electronics /Diode Circuits<br />

current to a safe value so that the Zener diode does not overheat Moreover, as will be shown<br />

later, the larger the value of R, the lower the output voltage variation for a given diode.<br />

'o<br />

Figure 3.28 A simple regulator circuit that provides a nearly constant<br />

output voltage from a variable input voltage<br />

Assuming that the characteristic of the diode is available, one can use a load-line construction<br />

(see Section 3.2) to analyze the operation of the circuit. As before, we use Kirchhoff s voltage law<br />

to write an equation relating v D to ijy. In this circuit, the diode operates in the breakdown region<br />

with negative values of v D and /p. For the circuit of Fig. 3.28, one obtains<br />

V ss + Ri D +v D =0 (3.31)<br />

This is the equation'of a straight line, so location of any two points is sufficient to construct the<br />

load line. Inspection of (3.31) shows that the slope of the load line is -1/R. Thus a change of the<br />

input voltage changes the position but not the slope of the load line. The intersection of the load<br />

line with the diode characteristic yields the operating point.<br />

Example 3.5 The voltage regulator circuit of Fig. 3.28 has R = 1 kfl and uses the Zener diode<br />

having the characteristic shown in Fig. 3.29. Find the output voltage for V ss — 15 V. Repeat for<br />

^=20V.<br />

Solution The load lines for both values of V ss are shown in Fig. 3.29: The output voltages are<br />

determined from the operating points where the load lines intersect the diode characteristic. The<br />

output voltages are found to be «y=10.0V for V SS =\5V and t^lO.SV for V SS =20V. Thus a 5 V<br />

change in the input voltage results in only a 0.5-V change in the regulated output voltage.<br />

irfmA)<br />

vjbOO<br />

Figure 3.29 Graphical analysis of the circuit of Fig. 3.28<br />

Actual Zener diodes are capable of much better performance than the one illustrated by Example<br />

3.5. The slope of the Zener diode characteristic has been accentuated in Fig. 3.29 for clarity.<br />

Actual Zener diodes have a more vertical slope in breakdown.<br />

The circuit of Fig. 3.28 is known as a shunt regulator because the Zener diode is connected in<br />

parallel (shunt with the load. The circuit is fed with a voltage that, as indicated in Fig. 3.27, is not<br />

very constant; it includes a large tipple component from a rectifier circuit. The load can be a<br />

simple resistor or a complex electronic circuit.


Analog Electronics / Diode Circuits<br />

The function of the regulator is to provide an output voltage v L that is as constant as possible in<br />

spite of the ripples in V ss and the variation in the load current I L . Two parameters can be used<br />

to measure how well the regulator is performing its function: the line regulation and the load<br />

regulation. The line regulation is defined as the change in V L corresponding to 1-V change in<br />

V ss<br />

Line regulation = AV,<br />

(3.32)<br />

AK<br />

ss<br />

and is usually expressed in mV/V. The load regulation is defined as the change in<br />

corresponding to a small change in I L<br />

Load regulation = —— (3.33)<br />

V L<br />

Variable<br />

supply<br />

V L =-v D<br />

0 (b)<br />

Figure 3.30 Piecewise linear model (a) for the Zener diode, valid for VD


Analog Electronics / Diode Circuits<br />

3.7 Wave-Shaping Circuits<br />

A wide variety of wave-shaping circuits find application in electronic systems. One of them is<br />

in function generators used to generate electrical test signals for laboratory work. In a function<br />

generator, a switching oscillator is used typically to generate a square wave. This square wave is<br />

then passed through a circuit that integrates it, resulting in a triangular waveform. Then the<br />

trianguk: waveform is passed through a carefully designed wave-shaping circuit to produce<br />

sinusoidal waveform. All three waveforms are available to the user. We will consider the design<br />

of such a function generator later. Numerous examples of wave-shaping circuits can be found in<br />

transmitters and receivers for television, as another example. In this section we discuss a few<br />

examples of wave-shaping circuits that can be constructed using diodes.<br />

Clipper Circuits. Diodes can be used to form clipper circuits in which a portion of input<br />

signals waveform is "clipped" off. For example, the circuit of Fig. 3.32a clips off any part of die<br />

input waveform above 6V and below -9V. (We are assuming ideal diodes.) If the input voltage is<br />

between -9V and 6V, both diodes are off and no current flows through them. Then there is no<br />

voltage drop across R and the output voltage v 0 is equal to the input voltage v s . On the other<br />

hand, if v s is larger than 6V, diode A. is on and the output voltage is 6V because the diode<br />

connects the 6-V battery to the output terminal. Similarly, if v s is lower than -9V, diode B is on<br />

and the output voltage is -9V. The output voltage resulting from 15V-peak triangular input is<br />

shown in Fig. 3.32b. The transfer characteristic of the circuit is shown in Fig. 3.32c.<br />

-73-


Analog Electronics /Diode Circuits<br />

Figure 3.32 Clipper circuit (a), its input and output waveforms (b) and<br />

transfer characteristic (c)<br />

The resistor R is selected large enough so that the forward diode current is within reasonable<br />

bounds (usually, a few milliamperes) but small enough so that the reverse diode current results in<br />

a negligible voltage drop. Often, we find that a wide range of resistance values provide<br />

satisfactory performance in a given circuit.<br />

In Fig. 3.32 we have assumed ideal diodes. If low-current silicon diodes are used, we expect a<br />

forward drop of about 0.6V, so we should reduce the battery voltage to compensate.<br />

Furthermore, batteries are not desirable for use in circuits if they can be avoided, because may<br />

need periodic replacement. Thus a better design uses Zener diodes instead of batteries. Practical<br />

circuits equivalent to Fig. 3.32a are shown in Fig. 3.33. 'Tie Zener diodes are labeled with their<br />

breakdown voltage. The circuits shown in Fig. 3.33 have learly the same performance as the<br />

circuits of Fig. 3.32.<br />

/f=2kii<br />

/f=2kn<br />

v o (0<br />

. 00 (b)<br />

Figure 3.33 Circuit of Fig. 3.32a with batteries replaced by Zener diodes and<br />

allowance made for a 0.6 forward diode drop (a); simpler circuit (b)<br />

Exercise 3.9 Sketch the transfer characteristic for the circuits of Fig. 3.34a and 3.34b. Allow for<br />

a 0.6V forward drop for die diodes. Sketch the output voltage if *> s (/)=15sin(ey/).<br />

-74-


Analog Electronics /Diode Circuits<br />

(a)<br />

Figure 3.34 Clipper circuits for Exercise 3.9<br />

(b)<br />

Clamp Circuits. Another diode wave-shaping circuit is the clamp circuit that is used to add a<br />

dc component to an ac-input waveform, so that the positive (or negative) peaks are forced to<br />

take a specified value. In other words, the peaks of the waveform are. "clamped" to a specified<br />

voltage value. A simple clamp circuit example is shown in Fig. 3.35. In this circuit, the positive<br />

peaks of die input are clamped to -5V. As an application example, clamp circuits are used to<br />

restore die dc component (black picture level) of video signals transmitted through ac-coupled<br />

channels.<br />

In the circuit of Fig. 3.35 the capacitor is a large value, so it discharges only very slowly and we<br />

can consider the voltage across the capacitor to be constant. Because the capacitance is large, die<br />

capacitor has very small impedance for the ac-input signal. Thus the output voltage of the circuit<br />

is given by<br />

v o (0 = v 5 (0-^c (3-35)<br />

If die positive swing of the input signal attempts to force the output voltage to become more<br />

positive than -5V, the diode conducts, charging die capacitor and thus increasing the value of V c .<br />

Thus the capacitor is charged to a value that adjusts the maximum positive value of the output<br />

voltage to -5V. A large resistor R is provided so that the capacitor can slowly discharge. This is<br />

necessary so that the circuit can adjust if the input waveform changes to smaller peak amplitude.<br />

Of course, one can change the voltage to which the circuit clamps by changing the battery<br />

voltage in Fig. 3.35. Reversing the direction of the diode causes the negative peak to be clamped<br />

instead of the positive peak. Switched, or pulse operating clamp circuits are actually used in the<br />

video e upment.<br />

v„(V)<br />

Figure 3.35 Example clamp circuit (a); output waveform for<br />

triangular, 0 dc, 10-V peak-to-peak input (b)<br />

Selection of R and C values for clamp circuits is a compromise. On one hand, we want die<br />

capacitor to have very small impedance compared to the resistor for the ac signal. This is<br />

necessary because we want the ac part of the output waveform to be identical to the input. On<br />

the other hand, if we make the RC time constant too long, the circuit takes a long time to adjust<br />

to changes in input amplitude. For now we can choose R to be a fairly large resistor, say about<br />

100 kQ, so diat the peak diode currents are not required to be large (more than a few<br />

milliamperes). Then we pick C so that the RC time constant is large compared to the period of<br />

die ac input signal, say by an order of magnitude. (In the case of video signal, the discharge<br />

-75-


Antdog Electronics / Diode Circuits<br />

period is equal to 64 |is, which is the duration of a single picture line.) This gives a clamp circuit<br />

with approximately the desired clamping action. Then we can simulate the circuit and adjust<br />

values until the performance is satisfactory. Finally, we can construct the circuit to measure its<br />

performance.<br />

3.8 Switching and High-Frequency Behavior of the pn Junction<br />

We have seen that the pn junction conducts little current when reverse biased and easily conducts<br />

a lot of current when forward biased. In many applications, such as high-speed logic circuits and<br />

high-frequency rectifiers, diodes that can switch rapidly between the conducting and<br />

nonconducting states are extremely desirable. Unfortunately, thepn junction displays two charge<br />

storage mechanisms that introduce delays and slow down the switching. Both of these<br />

mechanisms can be modeled as nonlinear capacitances.<br />

Due to the presence of the charge-storage effects, one has to distinguish between static and<br />

dynamic properties of any electronic device, using the diode as the simple example. The diode's<br />

familiar i-v characteristic is a static characteristic. For a one-port the word "static" implies that<br />

operation is described by an algebraic equation, i D —fiv^j, that relates corresponding dc voltage and<br />

current values, this is Equation (3.1) in the case of ideal pn junction. A device's static<br />

characteristic also relates time-varying voltages and currents, but only if the time variations are<br />

not "too fast". Electronic devices always contain internal capacitances that modify device<br />

behavior for fast signals. Therefore, these devices are described by differential equations, which<br />

reduce to algebraic equations for sufficiently slow signals. Thus "static" means that the device<br />

variables are changing at such low rates that their time derivatives in the differential equation are<br />

small enough to ignore. At high frequencies, derivatives of the circuit variables cannot be<br />

neglected and the device capacitances significantly affect the circuit performance. Before we<br />

consider charge storage in pn junctions, we briefly review conventional linear capacitors.<br />

A capacitor is constructed by separating two conducting plates by an insulator as shown in Figure<br />

3.36a. If voltage is applied to the capacitor terminals, charge flows in and collects on one plate.<br />

Meanwhile, current flows out of the other terminal and a charge of opposite polarity collects on<br />

the other plate. Positive charge accumulates on the plate to which the positive voltage is applied.<br />

This is illustrated in Figure 3.36b.<br />

Q<br />

Conducting<br />

plates<br />

Insulating<br />

dielectric ++++++++++++<br />

Figure 3.36 Parallel-plate capacitor (a), applying voltage to a capacitor causes a charge +J2 to accumulate<br />

on one plate and -Q to accumulate on the other plate (b)<br />

The magnitude of the net charge Q on one plate is proportional to the applied voltage V. Thus<br />

we have<br />

Q=CV (3.36)<br />

For a parallel-plate capacitor such as that shown in Figure 3.36, the capacitance is given by<br />

-76-


Analog Electronics / Diode Circuits<br />

-f<br />

(3.37)<br />

where A is the area of one plate, d is the distance between the plates, and e is the dielectric<br />

constant of the material between the plates. Often, the dielectric constant is expressed as<br />

e = e r e 0 (3.38)<br />

where e r is the relative dielectric constant and Eo = 8.85xl0" 12 F/m'is the dielectric constant for<br />

vacuum. [Actually, (3.37) is an approximation valid for d much smaller than both the length and<br />

the width of the plates.] Notice that the capacitance of the parallel-plate capacitor is proportional<br />

to the area of the plates and inversely proportional to the distance between the plates.<br />

Depletion Capacitance. Now consider ike. pn junction under reverse bias. As the magnitude of<br />

the voltage applied to the junction is increased, the field in the depletion region becomes<br />

stronger, and the majority carriers are pulled back farther from the junction. This is illustrated in<br />

Figure 3.37.<br />

The charge in the depletion region is similar to the charge stored on parallel-plate capacitor.<br />

Unlike the parallel-plate capacitor, a larger distance separates each additional increment of charge<br />

stored in the depletion region. Thus the reverse-biased junction behaves as a capacitor, but its<br />

equivalent plates move apart with the voltage, so the capacitance is not constant. The stored<br />

charge is not proportional to the applied voltage. This capacitance is called the depletion<br />

capacitance. Because the relationship between the stored charge and the applied voltage is not<br />

linear, we say that the depletion capacitance is nonlinear.<br />

C j = dv D<br />

(3.39)<br />

in which dQ is the differential of the charge stored in one side of the depletion region, and dv D is<br />

the increment in the voltage which caused the charge increment. C\ is the capacitance of the<br />

diode for a small ac signal superimposed on a dc operating point.<br />

It can be shown that the incremental depletion capacitance is given by<br />

BIBLIOTEKA GLOWNA PL<br />

QWb^ 5<br />

-77


Analog Electronics / Diode Circuits<br />

C J=-<br />

1-<br />

V J0]<br />

~\m<br />

(3.40)<br />

in which Cjo is the incremental depletion capacitance for zero bias, v D is the voltage across the<br />

diode (which is negative for reverse bias), V® is the built-in barrier potential (typically about 1 V),<br />

and m is called the grading coefficient. For a linearly graded junction m — 1 /3, and for an abrupt<br />

junction m— 1/2.<br />

The zero-bias depletion capacitance Cjo is approximately proportional to the area of the junction.<br />

Thus it is larger for high-power rectifiers, which must be physically large to accommodate high<br />

power dissipation. The value of Cjo also depends on doping levels. In highly doped junctions, a<br />

large amount of charge can be stored close to the junction - similar to the parallel-plate capacitor<br />

with small plate separation. Thus we find high values of CJO for highly doped junctions and low<br />

values for lighdy doped junctions.<br />

A reverse-biased diode can be used in circuits as a variable, voltage-controlled capacitor. A<br />

control signal reverse biases the diode, and as the control signal changes, the capacitance of the<br />

diode varies accordingly. -Diodes designed to have smoothly varying capacitance over a wide<br />

range of control voltage are called variable-capacitance diodes or varicaps. They can be used<br />

in. an JLC resonant circuit to vary the resonant frequency. Variable capacitance diodes can be used<br />

to design bandpass filters for which a control filter vary the center frequency, e.g. to tune in to an<br />

AM radio station. Varicaps are also useful in the design of voltage-controlled oscillators in which<br />

the frequency of oscillation depends on the diode capacitance. An application of this is the<br />

automatic frequency control (AFC) circuit of an FM radio. Manufacturers offer diodes intended<br />

for these applications having zero-bias capacitance Cp ranging from 10 to 1000 pF.<br />

For the MV2201 variable-capacitance diode (Motorola), approximate values of the parameters of<br />

formula (3.40) are CJO = 15 pF, m = 0.43, and KJO = 0.75 V. A plot of the depletion capacitance<br />

versus bias voltage using these parameters is shown in Figure 3.38.<br />

Figure 3.38 Depletion capacitance versus bias voltage for the MV2201 varicap diode<br />

Diffusion capacitance. Another basic charge-storage mechanism occurs when thepn junction is<br />

forward biased. For simplicity, we consider an abrupt junction with much heavier doping on the<br />

p-side than on the »-side (Le. N A »N^). Sometimes this is called a/> + » junction, where "plus"<br />

refers to heavy doping of the />-material. For such a diode under forward bias, die current<br />

-78-


Analog Electronics /Diode Circuits<br />

crossing the junction is due mainly to holes crossing from the/>-side to die /x-side. The current of<br />

electrons that are injected from the /r-side to the/>-side can be neglected.<br />

P<br />

P<br />

n-type<br />

n-type<br />

Stored<br />

charge<br />

Stored<br />

charge<br />

(a)/ D =J,<br />

(b)/ D =^>I,<br />

Figure 3.39 Hole concentration versus distance for two values of forward current<br />

Consider the hole concentration of the forward-biased p*n junction shown in Figure 3.39. The<br />

charge associated with the' holes that have crossed the junction is stored charge and is<br />

represented by the areas indicated in the figure. The charge is stored because a finite time passes<br />

before an average hole disappears due to electron-hole recombination. This time is called<br />

minority carriers lifetime. As the forward current is increased, more holes cross the junction and<br />

the stored charge increases. Because this charge is associated with bain that ice diffusing across<br />

the junction, we call the effect diffusion capacitance.<br />

It can be shown that the incremental diffusion capacitance is given approximately by<br />

in which T T is a parameter known as the transit time of the minority carriers. For the p + n<br />

junction, x T — Z" p is the lifetime of the holes on the »-side of the junction. On the other hand, f *. r<br />

the n + p junction, we have T T = r n , which is the lifetime of the free electrons on the/>-side. For \<br />

junction with comparable doping levels, r T is the weighted average of both lifetimes. Finally, I 0 s<br />

the j2-point diode current, and V T is the thermal voltage.<br />

9.2V It<br />

JT) c 40*B-9»IMlote><br />

Figure 3.40 Diffusion capacitance versus voltage for the 1N4148 diode<br />

Notice that the diffusion capacitance is proportional to the diode current. Thus the diffusion<br />

capacitance, like the current, increases rapidly when the voltage V D exceeds approximately 0.6 V<br />

for silicon devices at room temperature. A plot of the diffusion capacitance- of the 1N4148<br />

switching diode is shown in Figure 3.40. Under forward bias conditions, the diffusion capacitance<br />

is much larger than the depletion capacitance. (For the 1N4148 device, the depletion-capacitance<br />

-79-


Analog Electronics /Diode Circuits<br />

parameter Cjo is approximately equal to 2 pF.) However, diffusion capacitance is negligible for<br />

reverse bias.<br />

Small-signal diode model at high frequencies. A small-signal equivalent circuit for the pnjunction<br />

diode is shown in Figure 3.41. The resistance R* represents the ohmic resistance of die<br />

material on both sides of the junction; r& is the dynamic resistance of die pn junction that is<br />

discussed in Section 3.4. Its value is given by Equation (3.20), which is repeated here for<br />

convenience<br />

nV T<br />

r *=-T (3-42)<br />

1 D<br />

Cj is the depletion capacitance and Cdif is the diffusion capacitance.<br />

All die equivalent circuit parameters except R» depend on the bias point Under reverse-bias<br />

conditions, Cdif is zero, and ra is an open circuit. Hence die equivalent circuit simplifies as shown<br />

in Figure 3.41b.<br />

This equivalent circuit is vahd for the ptt junction over a wide range of frequencies, provided that<br />

small-signal conditions apply. At very high frequencies, lead inductances and stray capacitances of<br />

the diode physical structure are added to provide adequate accuracy.<br />

Diodes are most often used with large signals, and their nonlinear behavior must be taken into<br />

account. Computer modeling does this most easily. We illustrate this widi a few examples.<br />

(a) Forward bias<br />

Figure 3.41 Small signal linear circuits for the/>«-junction diode at high frequencies<br />

Large-Signal Diode Switching. Consider the circuit shown in Figure 3.42. The waveform of<br />

die source voltage t>s is shown in Figure 3.43a. Until / = 10 ns, v, is +50 V and the diode is<br />

forward biased. At / = 10 ns, the source voltage jumps rapidly to -50 V, reverse biasing the diode.<br />

rj] tf=5kO ryj<br />

Figure 3.42 Circuit illustrating switching behavior of a/>«-junction diode<br />

-80-


Analog Electronics /Diode<br />

Circuits<br />

-81-


Analog Electronics /Diode Circuits<br />

(d) Voltage across the diode (expanded scale)<br />

Figure 3.43 Waveforms for the circuit of Figure 3.42<br />

The resulting diode current is shown in Fig. 3.43b. As one might expect, the diode current is<br />

approximately (50 V)/(5 \£l) = +10 mA until / = 10 ns. Then the source voltage jumps to -50 V.<br />

Instead of dropping immediately to zero the diode current reverses to I R = -10 mA. At<br />

approximately / = 18.3 ns, -the current begins to fall in magnitude and approaches zero at / = 25<br />

ns. In the interval immediately after the source reverses polarity, the diode continues to act as it<br />

were forward biased. This is called the storage interval ft, as labeled in Figure 3.43.<br />

We can explain the behavior of the diode as follows. (To simplify the discussion, we assume a<br />

diode that is heavily doped on the/>-side compared to the «-side.) When forward bias is applied,<br />

holes flow across the junction into the »-side. The holes are minority carries there that diffuse<br />

into the «-side and eventually recombine with free electrons. When tfe reverses polarity, the holes<br />

stored in the »-side can again cross the junction back to the />-side. Until the supply of excess<br />

holes on the «-side is exhausted, current can easily flow in the reverse direction. This explains the<br />

storage interval of the diode current waveform.<br />

It can be shown that the storage interval for a/>»-junction diode is given by<br />

*,Z (3.43)<br />

B(l + B)<br />

in which T T is the transit time for the minority carriers and B= \I R /I F \ • I F is the forward current<br />

before switching and J R is the reverse current during the storage interval. The storage time<br />

becomes shorter at larger ratios B.<br />

After the excess holes have all recrossed the junction (or recombined with free electrons), the<br />

depletion capacitance of the diode is charged through the resistor. Thus, after the storage<br />

interval, we see an approximate exponential transient for the current in Figure 3.43b. (Since the<br />

depletion capacitance is nonlinear, the transient is not precisely exponential, as would be the<br />

transient in a linear RC circuit.) The interval for this transient, called transition time, is denoted<br />

by ft. By definition, the end of the transition interval occurs when the reverse diode current has<br />

reached a specific value, typically I R /10.<br />

The total time interval for the diode to become an approximate open circuit, called the reverse<br />

recovery time, is denoted by ftr. It is the sum of the storage time and the transition time.<br />

'«•='*+'/ (3-44)<br />

The transition time ft depends on the circuit resistance. Even though the depletion capacitance is<br />

nonlinear, the transition time is proportional to the circuit resistance.<br />

-82-


Analog Electronics /Diode Circuits<br />

The diode voltage is shown in Figure 3.43c. To be able to see the details of the diode voltage<br />

during forward bias, an expanded-scale waveform is also plotted in Figure 3.43d. Notice that the<br />

diode voltage remains positive during the storage time, showing that the diode continues to act as<br />

if it were forward biased, even though the current is in the reverse direction. The terminal voltage<br />

of the diode does fall somewhat when the current reverses, because the voltage drop across the<br />

ohmic resistance R, of the diode reverses polarity. Prior to / = 10 ns, the terminal voltage is the<br />

sum of the junction voltage and the ohmic drop. On the other hand, between / = 10 ns and t -<br />

18.3 ns the terminal voltage is the junction voltage minus the ohmic drop.<br />

Exercise 3.10 Consider the parallel-plate capacitor shown in Figure 3.36. The plates have<br />

dimensions of 20 |wm x 30 \ltn. (These dimensions are typical of the area of an integrated-circuit<br />

/>»-junction diode.) The relative dielectric constant of the material is ft = 11.9. (This is the value<br />

for silicon.) The capacitance is 1 pF, which is a typical zero-bias depletion capacitance for a lowpower<br />

diode. Find the distance between the plates. (The answer is the approximate zero-bias<br />

thickness of the depletion region.) Ans. d= 6.32x10 -8 m.<br />

Exercise 3.11 A certain abrupt-junction diode has a zero-bias depletion capacitance of 5 pF and<br />

a built-in barrier potential of 0.8 V. Compute the depletion capacitance for a reverse-bias voltage<br />

of (a) 5 V, (b) 50 V. Ans. 1.86 pF, 0.627 pF.<br />

Exercise 3.12 A certain diode has a transit time of 10 ns. Find values for the small-signal<br />

resistance and diffusion capacitance at I D = 5 mA. Assume the emission coefficient n-\ and a<br />

temperature of 300 K. Ans. r& - 5.2 Q., Cm = 1920 pF.<br />

Exercise 3.13 Consider the circuit of 3.42 with R changed to 50 kfi and with the source<br />

waveform of Figure 3.43a. (a) Think about the circuit and sketch the current versus time. Make<br />

use of results shown in Figure 3.43, use (3.43) to estimate / r Like the time constant of an RC<br />

circuit, /, is approximately proportional to R. (b) Write a PSpice program to obtain the plot of the<br />

diode current. Compare the results. Ans. The current waveforms are similar to Figure 3.43b with<br />

I F = 1 mA and I R = -1 mA, 4 = 8ns and A = 50 ns.<br />

Exercit : 3.14 Consider the circuit of 3.42 with the input voltage "being equal to -50 V in die<br />

time interval from 0 to 10 ns, then jumping up to reach 50 V at 10.01 ns, staying at 50V until 30<br />

ns and returning to -50 V at 30.01 ns. Write a PSpice program to obtain the plot of the diode<br />

current. Explain.<br />

TABLE 3.1 SPICE PARAMETERS FOR INJUNCTION DIODES<br />

Text<br />

notation<br />

SPICE<br />

notation<br />

Default<br />

value<br />

I IS 1.0E-14A<br />

n N 1<br />

V, BV oo V<br />

h IBV 1.0E-3A<br />

R, RS 0Q<br />

c* CJO OFyfV<br />

m M 0.5<br />

y t V] 1.0 V<br />

b r r TT 0s<br />

-83-


Analog Electronics / Diode Circuits<br />

SPICE ParatL. jis for Diodes. Figure 3.41a is the SPICE equivalent circuit for th&pn junction<br />

diode. Table 3.1 lists nine parameters required to model the static and dynamic behavior of the<br />

diode. The first four, describe the static properties of the diode, as expressed by the Schockley<br />

equation (IS, N) and the following equation that approximates the breakdown component of the<br />

diode current<br />

i D = -I z exp - V Z+v D<br />

(3.45)<br />

in terms of parameters (VB and IBV), where VB = V z is the breakdown voltage and IBV = I z is<br />

die current that flows through the diode at v D = V z . The remaining five parameters (RS, CJO, M,<br />

YJ, TT) describe the dynamic properties of the diode, following Equation (3.40) for depletion<br />

capacitance and (3.41) for diffusion capacitance. The default values of the two key dynamic<br />

parameters, CJO and TT, are both zero. This means that tbcdefault diode model in SPICE is a static<br />

model. To simulate dynamic effects, these default values must be overridden by nonzero values in<br />

.MODEL statement. An example diode model for the 1N4148 low-power general-purpose diode<br />

can be specified as follows:<br />

.model D1N4148 D(I»-0.1p Rs=16 CJO=*2p Tt=12n Bv«100 Ibv=0.1p)<br />

3.9 Special Diodes<br />

3.9.1 Schottky diodes<br />

The junction between certain metals (e.g. aluminum) and lightly doped n-type material forms a<br />

Schottky diode. A typical structure is shown in Figure 3.44. (On the other hand, a metallic<br />

contact with heavily doped »-type material results in an ohmic contact.)<br />

The detailed theory of the Schottky diode is somewhat different from that of the pn junction.<br />

However, the form of the results is the same - the current in the Schottky diode is given by the<br />

Shockley equation. The saturation current 7, is much higher for Schottky diodes than for pn<br />

junctions of the same size. A typical value for a Schottky diode is 7 S = 10" 10 A, whereas for a<br />

typical IC pn junction it is 7 S = 10" 16 A. Because of the larper value of 7 S , the forward voltage of<br />

the Schottky diode is significandy smaller (about 0.4 V compared to 0.7 V for a silicon pn<br />

junction diode).<br />

Schottky j— Ohmic<br />

contact -v / contact<br />

Figure 4.44 Schottky diode<br />

Another important difference is that the Schottky diode does not display charge storage when<br />

being switched from forward conduction to reverse-bias conditions. This is because for forward<br />

bias electrons rather than holes are injected from metal anode into the n material. Thus majority<br />

carriers (i.e. electrons) carry the current, and storage of minority carriers does not occur. Hence<br />

switching tends to be faster for the Schottky diode. This fact can be used to advantage in die<br />

design of fast logic gates. An example of a bipolar transistor inverter with a Schottky diode is<br />

discussed in Chapter 6.<br />

-84-


Analog Electronics/Diode Circuits<br />

3.9.2 Light-Emitting Diodes and Photodiodes<br />

Light emitting diodes (LEDs) are junctions constructed of special semiconductor materials<br />

(compounds) such as gallium arsenide-phosphide. In these devices,' the injected minority carriers<br />

that result from forward biasing give up energy in the form of radiated light when they<br />

recombine. Wavelength ranging from ultraviolet through the visible and into die infrared bands is<br />

obtainable by use of different semiconductors and doping impurities. LEDs are widely used as<br />

display devices, for transferring information into optical fibers, and for optical isolation as<br />

discussed shortly<br />

[••<br />

'D<br />

+<br />

0.7V<br />

V D<br />

Figure 3.45 Volt-ampere curves of photodiode/solar cell<br />

A photodiode is a junction in which photons of energy in incident light break covalent bonds,<br />

adding the drift of these new carriers to the existing reverse saturation current. Figure 3.45 shows<br />

die volt-ampere curve of such a diode as a function of increasing incident light intensity. In tiie<br />

third and fourth quadrants the device functions as a (dependent) current source controlled by<br />

light. I s is the amount of current that can flows through the diode in dark, at a given temperature<br />

Operation in the third quadrant represents passive conversion of light intensity information in<br />

electrical information. This is the photodiode functioning as an optical to electrical transduce-<br />

Applications include light meters and communication systems that receive information coded in<br />

the form of light. The obvious circuit model is a light-intensity-controlled current source.<br />

Diodes designed for fourth quadrant operation are called photovoltaic cells or solar cells.<br />

Points on the curves in the fourth quadrant correspond to voltage and currents of opposite sign,<br />

implying an active device that delivers power to an external circuit. The curves emphasize that the<br />

solar cell is an active device for voltages not exceeding 0.7 V. Series connection of many solar<br />

cells produce large dc voltages. Parallel connection of solar cells of large area generates high<br />

output currents. Thus series-parallel interconnections can convert relatively large amounts of<br />

solar power into electrical power at convenient voltage levels. Solar cell efficiency (electrical<br />

output power divided by solar input power) is typically of the order of 10 -15 %.<br />

Figure 3.46a shows how the light-emitting diode and the photodiode are combined in a useful<br />

device called an optical isolator. The signal th(t) is transferred to the load by means of light, with<br />

no physical connection whatsoever between the input and output circuits. The optical isolator is<br />

useful in many applications, including computer interfaces for biomedical instrumentation. In<br />

these applications, we want to transmit information while protecting a human subject and<br />

delicate equipment from dangers imposed by high voltages in the input circuit Figure 3.46b<br />

illustrates the principle. Notice that the hazard voltage that exists in the circuit on the left is not<br />

transferred to the load circuit on the right. Transformers also give isolation; however, die optical<br />

isolator operates at frequencies down to dc and is smaller, lighter and less expensive.<br />

-85-


Analog Electronics / Diode Circuits<br />

3.9 Summary<br />

The semiconductor diode is basically a pn junction. Such a junction is formed in a single<br />

semiconductor material crystal, most often silicon crystal. In the forward direction, the ideal<br />

diode conducts any current forced by the external circuit while displaying a zero voltage drop.<br />

The ideal diode does not conduct in the reverse direction; any applied voltage appears as reverse<br />

bias across the diode. The unidirectional current flow property makes the diode useful in the<br />

design of rectifier circuits. The diode i-v characteristic is described by Schockley equation; the<br />

forward conduction of practical silicon diodes is accurately described by the relationship i =<br />

7sexp[f/(«l/ T )]. A silicon diode conducts a negligible current until the forward voltage is at least<br />

0.5 V. Then the current increases rapidly, with the voltage drop increasing by 60 or 120 mV<br />

(depending on the value of ») for every decade of current change. In the reverse direction, a<br />

silicon diode conducts a current of the order of 1 nA. This current is much greater than 7s and<br />

increases with the magnitude of reverse voltage. The diode i-v characteristic and parameters<br />

depend on temperature. Diodes designed to operate in the breakdown region are called Zener<br />

diodes. They are employed in the design of voltage regulators whose function is to provide a<br />

constant dc voltage tfiat varies litde with variation in power supply voltage and/or load current.<br />

A hierarchy of diode models exists, with the selection of an appropriate model dictated by the<br />

application. Analysis of diode circuits depends on viewing the diode as a 3-state device that, at a<br />

particular time operates in its ON, OFF or breakdown state. For each state, we replace the diode<br />

by a simple equivalent that approximates the diode's i-v characteristic and we then solve the<br />

resulting linear circuit to check our state assumptions. In many applications, a conducting diode<br />

is modeled as having a constant voltage drop, usually about 0.7 V. A diode biased to operate at a<br />

dc current I D has a small-signal resistance ra = nV T /I D . With the diode modeling and analysis<br />

techniques, we can describe many useful circuits such as half- and full-wave rectifiers, limiters,<br />

clamping circuits, voltage doublers and voltage regulators. These perform a variety of useful<br />

functions such as turning ac into dc, shifting waveform levels, modifying the shapes of time-<br />

-86-


Analog Electronics /Diode Circuits<br />

varying waveforms, and producing dc that is relatively independent of changes in loading and<br />

source voltage. There are two sources of parasitic capacitance within the pn junction: depletion<br />

capacitance, associated with the layers of bound ions in the depletion region, and diffusion<br />

capacitance, associated with storing excess minority charge carriers just outside the depletion<br />

region. When signals change slowly in diode circuits, these capacitances are insignificant, and the<br />

diode satisfies its static equation (i.e. Schockley equation). For rapid.transitions or fast signals,<br />

however, a more appropriate diode description is needed, which is a differential equation that is<br />

nonlinear when posed in terms of diode current or voltage. Because of diode capacitances, delays<br />

occur in diode switching, and waveform processing predicted upon static diode theory<br />

deteriorates at high frequencies. SPICE models for diodes were introduced. By understanding<br />

how SPICE parameters relate to the diode parameters we use in circuit analysis, we can use<br />

simulation to study with increased accuracy and precision both static and dynamic operation of<br />

diodes. With SPICE we can formulate more realistic circuit descriptions, relying on computation<br />

to verify and reinforce our understanding of diode circuits. Accuracy of computer simulation,<br />

such as obtained from SPICE, is limited by the approximations we made when building models<br />

for real-world, physical devices. A Schottky diode is based on metal-semiconductor junction. It<br />

features fast switching and lower voltage drop when forward biased as compared to pn junction<br />

diode. Light-emitting diode, the photodiode and optical isolator are examples of optoelectronic<br />

devices useful in telecommunications and instrumentation technology.<br />

Notes<br />

-87-


Analog Electronics / FET Circuits<br />

4 Field-Effect Transistor Circuits<br />

Now we turn our attention to devices that can amplify an input signal. In this chapter we<br />

consider the field-effect transistor (FET). In the first few sections of this chapter we describe<br />

the external characteristics of several types of FETs. Then we consider some simple but<br />

important circuits that use FETs, useful in amplifiers and logic circuits.<br />

4.1 The n-Channel Junction FET<br />

The simplified physical structure of an ^-channel junction field-effect transistor (FET) is<br />

shown in Fig. 4.1a, and the circuit symbol is shown in Fig. 4.1b. The device consists of a channel<br />

of »-type semiconductor with ohmic (nonrectifying) contacts at each end. These contacts are<br />

called the drain and the source. Alongside the channel, there are regions of ^-type<br />

semiconductor electrically connected to each other and to the gate terminal.<br />

(a)<br />

Figure 4.1 Simplified physical structure (a) of an ^-channel FET; its circuit symbol (b)<br />

The /w»-junction between the gate and the channel is a rectifying contact similar to />»-junction<br />

diodes discussed in Chapter 3. In almost all applications, this junction is reverse biased, so<br />

virtually no current flows in the gate terminal. Hence the gate is negative with respect to the<br />

channel in normal operation of an »-channel FET. (Recall that the/>-side is negative with respect<br />

to »-side for reverse bias of a/>»-junction.)<br />

Applying reverse bias voltage between gate and channel causes the depletion layer of the gatechannel<br />

junction to become wider. The greater the reverse bias, the wider this nonconductive<br />

layer becomes. Eventually, for VQS


Analog Electronics / FET Circuits<br />

'D<br />

Figure 4.2 Circuit for discussion of drain characteristics of the ^-channel FET<br />

Suppose that v GS is zero. Then, as v DS increases, i D increases as shown in Fig. 4.3. The channel is<br />

a bar of conductive material with ohmic contacts at the ends - exactiy the type of construction<br />

used for ordinary resistors. Therefore, it is not surprising the i D is proportional to v DSi for small<br />

values of t> DS .<br />

VG5=0<br />

Figure 4.3 Drain current versus drain-to-source voltage for<br />

zero gate-to-source voltage<br />

However, for larger values of v DS , drain current increases more and more slowly. This is because<br />

the end of the channel closest to the drain is reverse biased by the (v DS + \ v GS \) = v DS voltage. As<br />

v DS increases, the depletion layer becomes wider, causing the channel to have higher resistance.<br />

After the pinch-off voltage is reached, the drain current becomes nearly constant for additional<br />

Cutoff v cs


Analog Electronics / FET Circuits<br />

value is under the control of v GS . If v GS is less than the pinch-off voltage, v GS V P (4-4)<br />

In the linear region, the drain current is given by<br />

i D =K[2(v as -V J .)v DS -vl s ] (4.5)<br />

The constant K has units of current per volt 2 . Study of this equation for a fixed value of v GS<br />

shows that it describes a parabola passing through the origin of the ip-Vps plane. Furthermore,<br />

the maximum of the parabola is on the boundary between the linear and saturation regions.<br />

Saturation region. An ^-channel FET is in the saturation region if<br />

v G5 > V, (4.6)<br />

and if<br />

V GD=( V GS- V DS)< V P (4.7)<br />

In the saturation region, the drain current is given by<br />

i D =K(v GS -V p f (4.8)<br />

A plot of Equation (4.8) is shown in Fig. 4.5. This plot represents the so-called transfer<br />

characteristic of the FET in saturation.<br />

The drain current in the saturation region for v GS =0 is denoted as I DSS and is usually specified on<br />

manufacturers' data sheets. Substituting v GS =0 into (4.8) we find that<br />

Ioss=KV^ (4.9)<br />

Solving for K, we have<br />

* = 7T (4-10)<br />

If values are given for I DSS and Vp, the static characteristics of the JFET can be plotted. Typical<br />

values for a small-signal »-channel JFET are: I DSS =5-rl0mA and K P =-3V.<br />

-90-


Analog Electronics/FET Circuits<br />

IDSS<br />

+ *GS<br />

Figure 4.5 Plot of fa versus VDS in the saturation region of the FET<br />

Breakdown. As we mentioned earlier, there are several effects not modeled by the device<br />

equations we have given. An example of one of these effects occurs if the reverse bias between<br />

gate and channel becomes too large - then the junction experiences breakdown and the drain<br />

current increases very rapidly. Usually, the greatest reverse bias is at the "drain end of the channel,<br />

so breakdown occurs when V^Q exceeds the breakdown voltage V B in magnitude. Because v DG =<br />

V DS ' V GS> breakdown takes place at smaller values of v DS as P GS takes values closer to pinch-off.<br />

This is illustrated in Fig. 4.6. We seldom operate FETs in breakdown; however, breakdown<br />

voltage value V B should be taken into design consideration to avoid the device entering into this<br />

operation region.<br />

4.2 Me il-Oxide-Semiconductor FETs<br />

Another important class of devices is the metal-oxide-semiconductor field-effect transistor<br />

(MOSFET). There are two types, known as depletion MOSFETs and enhancement<br />

MOSFETs. Each of them can be realized as an w-channel or a/(-channel device. Basically, all of<br />

these FETs have very similar characteristics. Once you master one type, such as the »-channel<br />

JFET, it is much easier to assimilate the relatively minor differences between them. Circuit<br />

symbols for the »-channel and for the />-channel MOSFET devices are shown in Fig. 4.7. The<br />

circuit symbols for/>-channel are the same as the circuit symbols for n-channel devices except for<br />

the directions of die arrowheads. The MOSFET terminal labeled as B denotes the substrate or<br />

the body contact.<br />

Drain current versus v GS in the saturation region for »-channel devices is shown in Fig. 4.8. The<br />

depletion MOSFET has output characteristics nearly identical to those of die JFET. The main<br />

difference between die »-channel JFET and the »-channel MOSFET is in the fact diat the<br />

MOSFET can be operated with positive values of v Gs . (This is usually not done with the JFET<br />

because it would result in forward bias of the gate-to-channel junction and the corresponding<br />

-91-


Analog Electronics/FET Circuits<br />

increase of gate current.) The equations we have given in Section 4.1 for the ^-channel JFET also<br />

apply for the //-channel depletion MOSFET.<br />

Source<br />

<br />

(h)<br />

Figure 4.7 //-channel FETs circuit symbols (a, b, c),/(-channel FETs circuit symbols (d, e, f),<br />

simplified physical structure of the //-channel enhancement MOSFET (g)<br />

and the //-channel depletion MOSFET (h)<br />

No current flows in the //-channel enhancement MOSFET for v GS less than a certain positive<br />

value known as the threshold voltage which is denoted by V A . The equations we have given in<br />

Section 4.1 for the //-channel JFET also apply for the //-channel enhancement MOSFET if the<br />

(positive) threshold voltage Vth replaces the (negative) pinch-off voltage Vp. In particular, the<br />

enhancement MOSFET is cutoff for v GS < V&.<br />

The characteristics of the /(-channel FETs are the same as for the respective »-channel devices<br />

except that voltage polarities and current directions are inverted. If we continue to reference the<br />

drain current into the drain, the algebraic signs of the currents and voltages must be inverted for<br />

/(-channel devices. As a consequence, for the /(-channel devices the drain current is negative and<br />

the pinch-off voltage V p is positive for the JFET and for the depletion MOSFET. The threshold<br />

voltage Vfo for the/(-channel enhancement MOSFET takes on negative values.<br />

Figure 4.8 Transfer characteristics for FETs:<br />

(a) //-channel JFET and depletion MOSFET, (b) //-channel enhancement MOSFET,<br />

(c) /(-channel JFET and depletion MOSFET, (d)/(-channel enhancement MOSFET.<br />

-92-


Analog Electronics / FET Circuits<br />

Gate protection. Because of their construction, MOSFETs have extremely high input<br />

impedance between gate and channel - in excess of 1000MQ. This high impedance is due to a<br />

thin insulating (silicon dioxide) layer placed between gate and channel. In handling these devices,<br />

it is easy to develop sialic electric voltages greater than the breakdown voltage of the gate<br />

insulation. Breakdown of the insulating layer is destructive, usually resulting in a short circuit<br />

between gate and chat<br />

To alleviate this problem, the gate terminals are usually protected by back-to-back Zener diodes<br />

as shown in Fig. 4.9. If the device is exposed to a static electric charge, one of the Zener diodes<br />

breaks down (depending on die voltage polarity, the other diode is on), providing a<br />

nondestructive discharge path. Usually, die diodes are fabricated on die same chip of<br />

semiconductor as the FET.<br />

Variation in FET parameters with temperature. It is well known mat the semiconductor<br />

material mobility decreases with temperature. Because of more frequent collisions of charge<br />

carriers with the rapidly moving lattice ions. It turns out that the constant K in (4.5) and (4.8),<br />

which are valid for both JFETs and MOSFETs (with V p = Kth where appropriate), is<br />

proportional to the carrier mobility. Consequendy, their transfer characteristics change as in Fig.<br />

4.9b and 4.9c. It follows that the output characteristics for both devices crowd more closely as in<br />

Fig. 4.9d, juat opposite of bipolar transistor characteristics that are considered in Chapter 5.<br />

The threshold and pinch-off voltages decrease by about 2 mV/°C with increases in temperature,<br />

causing the curves in Fig. 4.9b and 4.9c to shift to the left with temperature increases as they<br />

droop downward.<br />

-93-


Analog Electronics / FET Circuits<br />

4.3 Load-Line Analysis of a Simple JFET Amplifier<br />

In this section we analyze the JFET amplifier circuit shown in Fig 4.10 by use of the graphical<br />

load-line approach. The batteries bias the JFET at a suitable operating point so that amplification<br />

of the input signal v s {f) can take place. We will see that the input voltage v s (t) causes v GS to vary<br />

with time, which in turn causes t D to vary. The changing voltage drop across R D causes an<br />

amplified version of the signal to appear at the drain terminal.<br />

Figure 4.10 Simple JFET amplifier circuit<br />

Applying Kirchhoff s voltage law to the input loop, we obtain the following expression:<br />

VGS(0 = V,(0-K GC (4.11)<br />

As an example, we assume that the input signal is a 1-V peak 1-kHz sinusoid and that V GG is IV.<br />

Thus we have<br />

v C5 (/) = sin(2000^) -1 (4.12)<br />

Writing a voltage equation around the drain circuit, we obtain<br />

V DD = R D i D {t) + v DS {t) (4.13)<br />

For our example, we assume that R D =l]sQ and V DD =20V, so Equation (4.13) becomes<br />

20 = * D (/) + v DS (0 (4.14)<br />

where we have assumed that i D (t) is in milliamperes. A plot of this equation on the drain<br />

characteristic of the transistor is a straight line called the load line. To establish the load line, we<br />

first locate two points on it. Assuming zero drain current, i D — 0, in (4.14) we find that v DS —2QW.<br />

These values plot as the lower right-hand end of the load line shown in Fig. 4.11. For a second<br />

point, we assume that v DS = 0, which yields / D =20mA when substituted into (4.14). This pair of<br />

values (y DS — 0 and i D = 20 mA) plots as the upper left-hand end of the load line.<br />

If v s {t) = 0, Equation (4.11) yields v GS = -VQQ — -1 V. Therefore, the intersection of the curve for<br />

P GS = -1- V with the load line is the quiescent operating point. The quiescent values are 7p = 9 mA<br />

andK D =llV.<br />

The maximum and minimum values of the gate-to-source voltage are J^cj-max = 0 V and V i GSnan<br />

= -2V [see Equation (4.12) and Figure 4.12]. The intersections of the corresponding curves with<br />

the load line are labeled as points A and 23, respectively, in Fig. 4.11. At point A, we find that<br />

^nrmin= 4 V and 1 ^ = 16 mA. At point B, we find that V DSm2x = 16 V and 1 ^ = 4 mA.<br />

94-


Analog Electronics / FET Circuits<br />

The plots of v s (t) and v DS (t) versus time are shown in Fig. 4.12. Notice that the peak-to-peak<br />

swing of the drain-to-source voltage is 12V, whereas the peak-to-peak swing of the input signal is<br />

2V. Furthermore, the ac voltage at the drain is inverted compared to the signal at the gate.<br />

Therefore, this is an inverting amplifier. Apparently, one can calculate the circuit gain as equal to<br />

A^ = -12/2 = -6, where the minus sign is due to the inversion.<br />

Figure 4.11 Drain characteristic and load line for the FET amplifier<br />

Notice, however, that the output waveform shown in Fig. 4.12 is not a symmetrical sinusoid Wte<br />

the input. For illustration, we see that starting from the j2-point at V D = 11 V, the output voltage<br />

swings down to Vi >m^a — 4 V for a change of 7 V. On the other hand, the output swings up to<br />

16V for a change of only 5V from the j2-point on the positive going half-cycle of the output We<br />

cannot properly define gain for the circuit because the ac output signal is not proportional to the<br />

ac input Apparendy the FET is a nonlinear device. Nevertheless, die output signal is larger th n<br />

die input signal even if it is distorted.<br />

The distortion is due to the fact that die characteristic curves for the FET are not uniformly<br />

spaced. Of course, if much smaller input signal were applied, we would obtain amplification<br />

without appreciable distortion.<br />

The rather modest gain (A^ = -6) that we see in this circuit is typical of RC-coupled FET<br />

amplifiers. In general, BJT amplifiers have much larger voltage gains. However, if we consider<br />

current gain, FET circuits have larger gains than most BJT circuits. For example, we have<br />

considered the input current for the circuit of Figure 4.10 to be zero. An infinite current gain<br />

would result if one attempts to calculate current gain as the ratio of the ac current in R D to tb-i<br />

input current. On die other hand, the current gain of a BJT ranges from about 10 to several<br />

hundred.<br />

The amplifier circuit we have analyzed in this section is fairly simple. Practical amplifier circuits<br />

are usually much more difficult to analyze by graphical mediods. Later in this chapter we develop<br />

a small-signal equivalent circuit for the FET, and then we can use mathematical circuit-analysis<br />

techniques instead of graphical analysis. Usually it is more useful for investigation of practical<br />

amplifier circuits. However, graphical analysis of simple circuits provides an excellent way to<br />

understand die basic concepts of amplifiers.<br />

-95-


Analog Electronics / FET Circuits<br />

4.4 The Self-Bias Circuit<br />

Analysis of amplifiers is undertaken in two steps. First, we analyze the dc circuit to determine the<br />

j2-point In this analysis, the nonlinear device equations and/or curves are used. Then, after the<br />

bias is found, we use a linear, small-signal equivalent circuits to find the input resistance, voltage<br />

gain, and so on. In this section and die next, we consider analysis and design of dc bias circuits<br />

for FETs.<br />

The two-battery bias circuit used in the amplifier of Fig. 4.10 is not practical. Usually one battery<br />

is available only. Even more significant problem is that FET parameters vary considerably from<br />

device to device. For a given type of JFET, I DSS may vary by a ratio of 5:1. Furthermore, the<br />

pinch-off voltage is different from device to device.<br />

Plots of i D versus v GS are shown in Fig. 4.13 for extreme FETs, all of which have die same<br />

manufacturer's type number. The range of variation shown is typical. Notice that if V GS were the<br />

same for all devices, a considerable variation in I D would occur. Some devices would be biased at<br />

one end of the load line and others at the opposite end. On the other hand, to obtain the<br />

maximum symmetrical swing of the output voltage without severe distortion, we require the<br />

operating point to be near the middle of the load line for all devices. Thus a fixed-bias circuit<br />

that maintains the same value of V GS independent of the device parameters is not suitable for<br />

mass production.<br />

Figure 13 Transfer characteristics of extreme devices for JFETs having the<br />

same type number. Bias with fixed VGS results in a large variation of h.<br />

-96-


Analog Electronics / FET Circuits<br />

A more practical circuit, known as the self-bias circuit, is shown in Fig. 4.14. The resistor R s is<br />

essential to the operation of this bias circuit. The resistor R^ is usually a large value (several<br />

megaohms) that maintains the dc voltage at the gate close to the ground and still provides high<br />

input impedance at the gate. Since only a small current (InA or less) flows in the gate, the dc<br />

voltage drop across RQ is negligible. In applications where input signal is not applied to gate, RQ<br />

can be replaced by a short circuit. The drain resistance R D is required to be present if we want an<br />

amplified signal to appear at the drain. However, in some circuits, the output is taken from the<br />

source terminal, and then we would replace R D by a short circuit.<br />

DD<br />

Figure 4.14 Self-bias circuit used for JFETs and depletion MOSFETs<br />

The drain current flows out through the source and through the source resistor R s , creating a<br />

voltage drop. Writing a voltage equation around the gate-source loop of Fig. 4.14 and neglecting<br />

the drop across RQ we have<br />

v G s=-Rsh (4-15)<br />

A plot of this equation is called the bias line and is shown in Fig. 4.15, which also shows i D<br />

versus v GS for extreme devices of a given type. The operating point is at the intersection of the<br />

bias line and the device transfer curve. Notice that V GS is smaller in magnitude for the lowcurrent<br />

device than for the high-current device. Thus the self-bias circuit adjusts V GS to<br />

compensate for changes in the device, thereby reducing variations in I D compared to a fixed-bias<br />

circuit.<br />

Bias line<br />

VQS=-RS'D<br />

Figure 4.15 Graphical analysis of the self-bias. The device-to-device variation of drain current is<br />

much less than for the fixed-bias circuit.<br />

The device curves shown in Fig. 4.15 are valid only if device operates in the saturation region.<br />

Usually, in FET amplifier circuits, operation in the saturation region is desired. However, we<br />

should check to make sure that V DS is large enough for operation in the saturation region before<br />

accepting results based on this analysis.<br />

VGS<br />

-97-


Analog Electronics / FET Circuits<br />

Exercise 4.1 Design a self-bias circuit for an »-channel FET having I DSS =4toA and V P =-2V.<br />

The circuit is to have R D =2.2kQ, V DD =20V, and Ij^ltnA. Use standard 10%-tolerance resistor<br />

values. Ans. Ry=270Q.<br />

Exercise 4.2 Analyze the self-bias circuit designed in Exercise 4.1. Repeat the analysis for a highcurrent<br />

device having 7 Dxr =8mA and V P =-4V. Ans. VQ^-OSGV, J D =2.07mA, V DS =U.9V;<br />

V GS =-l.l2V,I D =4.UmA, V DS =9.11V.<br />

4.5 The Fixed- plus Self-Bias Circuit<br />

The self-bias circuit gives fair performance in maintaining a fixed I D from device to device, but<br />

sometimes better performance is needed. The fixed- plus self-bias circuit shown in Fig. 4.16<br />

provides a solution.<br />

For purposes of analysis, we replace the gate bias circuit with its Thevenin equivalent, as shown<br />

in Fig. 4.16b. The Thevenin voltage is<br />

VQ<br />

r .=V n<br />

X<br />

DD R,+R,<br />

(4.16)<br />

and the Thevenin resistance RQ is the parallel combination of R f and R 2 . Writing a voltage<br />

equation around the gate loop of Fig. 4.16b, we obtain<br />

V G = v GS + R s i D (4.17)<br />

Notice &at we have assumed that the voltage drop across R^ is zero. Now, if we assume that the<br />

transistor is in the saturation, we have<br />

i D =K(v GS -V P f (4.18)<br />

Simultaneous solution of (4.17) and (4.18) yields the operating point (provided that it falls in the<br />

saturation region). Then we can find v DS by writing a voltage equation around the drain loop<br />

v DS =r D0 -(tf 0 + * s )/, (4.19)<br />

(a)<br />

(b)<br />

Figure 4.16 Fixed- plus self-bias circuit. Original circuit (a),<br />

Gate t>ias circuit replaced by its Thevenin equivalent (b).<br />

Figure 4.17 shows the graphical solution of Equations (4.17) and (4.18). Notice that higher values<br />

of V G result in smaller variation in I D for extreme devices because die bias line becomes closer to<br />

horizontal. (Also notice that V G =0 corresponds to the self-bias circuit.) However we must not<br />

-98-


Analog Electronics / FET Circuits<br />

choose V G too high because this raises the voltage drop across R s and sufficient voltage must be<br />

allocated for v DS and R D .<br />

Exercise 4.3 Design for V^SV a fixed- plus self-bias circuit for the JFET of Exercise 4.1<br />

having I DSS =4mA and V P =-ZV. The circuit is to have R p =2.2kQ, V DD =20V, and I^lmA. Use<br />

standard 10%-tolerance resistor values. Ans. Rj=2.7kfi.<br />

Exercise 4.4 Analyze the self-bias circuit designed in Exercise 4.3. Repeat the analysis for a highcurrent<br />

device having I DSS =SmA and V P =-4V. Ans. F GJ -=-0.564V, 7 D =2.06mA, V DS =9.9V;<br />

V GS =-\.76V,I D =2.50mA, V DS =7.73V.<br />

Figure 4.17 Graphical solution for the fixed- plus self-bias circuit.<br />

Note that ID is nearly independent of the device if VG yz large.<br />

Another advantage of the fixed- plus self-bias circuit is that it also works for enhancement<br />

MOSFETs. On the other hand, the self-bias circuit is not suitable for enhancement MOSFETs<br />

because the gate must be more positive than the source (assuming »-channel devices), which is<br />

not possible in the self-bias circuit.<br />

Example 4.1 Design a bias circuit for an «-channel enhancement MOSFET having K /4 =4V and<br />

K=10- 3 A/V 2 . The power supply voltage is V DD =\5V. A jg-point of ,K DJ -=5V and / p =5mA is<br />

desired. The circuit is to be used in an amplifier with the ac output taken from the drain terminal.<br />

Solution. Since an enhancement device is specified, we must use the fixed- plus self-bias circuit<br />

because the self-bias circuit would give a zero drain current regardless of the resistor value<br />

selected. The circuit is shown below. Values must be selected for R,, R^, R D and R s .<br />

Figure 4.18 Circuit of Example 4.1<br />

Since the supply voltage is 15V and a bias value V DS =5V is specified, a total of 10V remains to<br />

be divided between R D and R s . The drain current flows through both resistors, so we have<br />

-99-


Analog Electronics / FET Circuits<br />

R S +R D = {yoD ~^DSr ID = 2^ • Equation (4.17) shows that larger voltages across R s result<br />

in higher values for V G . As we have seen, large values of V G lead to good bias stability. Thus we<br />

are led to allocate a large portion of the available voltage to the drop across R s . In the extreme<br />

case, this would produce R D =0 and Ry=2k£2. However, since an ac output signal is to be<br />

developed at the drain, R D =0 is not acceptable. Thus we select R D =Rj-=lkfi. (This choice is<br />

somewhat arbitrary - perhaps in a complete amplifier design, the ac performance required would<br />

dictate particular choice.)<br />

Assuming the MOSFET is operating in the saturation region, we have<br />

Io = K{Vcs-V lh ) 2<br />

Substituting values and solving, we find that V GS = 6.24V. Thus the gate voltage is given by<br />

V G =V GS + R s I D =n.2V<br />

Now we must find values for R, and R^, so that the 15-V supply voltage divides with 11.2V<br />

across R^ and 15-11.2=3.8V across R ; . If we make somewhat arbitrary choice that<br />

R,+R 2 =1.5MQ, then R,=380kQ and R 2 =\A2MQ provide the desired voltage division.<br />

Therefore we choose the closest standard values, namely R,=390kfi and R 2 =1.1MQ<br />

Exercise 4.5 (a) Analyze jhe circuit designed in Example 4.1 to find of V DS and I D . The answer<br />

should verify that the operating point achieved is close to the design objectives, (b) Repeat the<br />

analysis for an «-channel enhancement MOSFET having V tb =5V and K=2taA/V z . Ans.<br />

7 D =4.87mA, K DJ =5.27V; J D =4.56mA, V DS =5.STV.<br />

Exercise 4.6 A certain «-channel enhancement MOSFET has V lt =2V and K=2mA/V 2<br />

The bias<br />

circuit (as shown in Fig. 4.18) has V DD =20V, R 7 =R 2 =1MQ, R D =lkO and Rj-=2.2kQ. Find V DS<br />

and I D . Ans. J D =3.07mA, V DS =\0.2V.<br />

Exercise 4.7 Find the largest value of R D that can be used in the circuit of Exercise 4.6 if the<br />

MOSFET is to remain in saturation. Assume that R , R 2 , and R s remain fixed. Ans.<br />

R Dmax =3.91kQ.<br />

4.6 The Small-Signal Equivalent Circuit<br />

In the preceding two sections, we considered dc bias circuits for FET amplifiers. Now we<br />

consider the relationships, between the signal currents and voltages for small changes from Q-<br />

point. As usual, we denote total quantities by lowercase letters with uppercase subscripts such as<br />

/ D (/) and v G5 (t). The dc j2~point values are denoted by uppercase letters with uppercase subscripts,<br />

such as J D and V GS . The signals are denoted by lowercase letters with lowercase subscripts such as<br />

*d(/) and %i(^. The total current or voltage is the sum of the j2-point value and the signal. Thus we<br />

can write<br />

i D (0 = I D +iAO (4-20)<br />

and<br />

v GS (0 = f GS +v ff (0 (4-21)<br />

In the following we assume that the FETs are biased in the saturation region, which is usually die<br />

case for amplifier circuits. Equation (4.8), repeated here for convenience, gives the total drain<br />

current in terms of the total gate-to-source voltage.<br />

-100-


Analog Electronics / FET Circuits<br />

i D =K(v GS -V P f (4.8)<br />

Substituting (4.21) and (4.20) into (4.8), we obtain<br />

I D +i d (0 = tfycs + v^ -V P ] 2 (4.22)<br />

The right-hand side of (4.22) can be expanded to obtain<br />

I D + /,(/) = K(V GS -V P f +2K(V GS -V P ) Vgs + Kv 2 m (4.23)<br />

However, the j2-point values are also related by (4.8), so we have<br />

Io=K{V GS -V P ) 2 (4.24)<br />

Therefore the first term on either side of Equation (4.23) can be cancelled. Furthermore, we are<br />

interested in small-signal conditions for which the last term on the right-hand side is negligible<br />

and can be dropped [i.e. we assume that |f G j(^| is much smaller than | Vcr^pW- With these<br />

changes, Equation (4.23) becomes<br />

(t) = 2K(v GS -V P )v gs (4.25)<br />

We define the transconductance of the FET as<br />

g„=2K(V GS -V P ) (4.26)<br />

Then Equation (4.25) can be written as<br />

'„(') = S.v,(0 ( 4 - 27 ><br />

The gate current of the FET is negligible, so we have<br />

/ f (0 = 0 (4.28)<br />

The small-signal equivalent circuit shown in Fig. 4.19 can represent equations (4.27) and (4.28).<br />

Thus die FET is modeled by a voltage-controlled current source connected between the drain<br />

and source terminals. The model has an open circuit between gate and source.<br />

Figure 4.19 Small-signal equivalent circuit for FETs<br />

Solving Equation (4.24) for the quantity (V GS - V P ) and substituting into Equation (4.26), we<br />

obtain<br />

g m = 24H~ D (4.29)<br />

Then if we use Equation (4.10) to substitute for K, we have<br />

which is often a convenient formula for computing the transconductance of a JFET or depletion<br />

MOSFET at a given j2-point. Of course, I DSS does not apply for enhancement MOSFETs, and<br />

Equation (4.29) applies for them.<br />

-101-


Analog Electronics / FET Circuits<br />

Figure 4.20 Small signal equivalent circuit that accounts for the dependence of ID on ws<br />

Furthermore, the first-order equations we have given for the FET do not include a term to<br />

account for the small effect of (&. on the drain current Previously, we assumed that the drain<br />

characteristics are horizontal in the saturation region, but this is not exact — the drain<br />

characteristics slope slightly upward with increasing Pd>. If we wish to account for the effect of &<br />

in the small-signal equivalent circuit, a resistance r& called the drain resistance is added between<br />

the drain and source as shown in Fig. 4.20. Equation (4.27) becomes<br />

U = Sm v ** + r ds<br />

(4.31)<br />

Figure 4.21 Determination of ra from the FET drain characteristics using (4.33)<br />

The definition of the reciprocal of n (which actually is the drain conductance gd=l/rd) is the<br />

partial derivative of drain current with respect to %., evaluated at j^-point:<br />

1 Si n<br />

(4.32)<br />

Sv DS<br />

-t*a<br />

Q-foial<br />

which can be approximated as<br />

1 A/ " (4-33)<br />

- uo<br />

where AI D is an increment of drain current centered at the j2-point. Similarly, Avds, is an<br />

increment of drain current centered at the jg-point. Equation (4.33) is useful to evaluate the drain<br />

resistance from the output characteristic of the FET plotted for a constant v GS =V Gs . This is<br />

illustrated in Fig. 4.21.<br />

Similarly, an alternative definition of the transconductance ^ is the partial derivative of I D with<br />

respect to v GS , evaluated at j2-point:<br />

on<br />

Sv, GS Q-poinl<br />

which can be approximated as<br />

(4.34)<br />

-102-


Analog Electronics / FET Circuits<br />

bm —<br />

ML<br />

Av<br />

GS<br />

(4.35)<br />

where At^ is an increment of drain current centered at die jg-point. Equation (4.35) is useful to<br />

evaluate die transconductance from die FET output characteristics plotted for two different<br />

values of gate-to-source voltage. This idea is illustrated in Fig. 4.22.<br />

Figure 4.22 Determination of g^ from die FET drain characteristics using (4.35)<br />

The device equations and the equivalent circuit that we have derived from them describe only the<br />

static behavior of die device. For rapidly changing currents and voltages, the additional<br />

capacitances are required for an accurate model. A small-signal FET model suitable to represent<br />

die JFET behavior for high-frequency operation is shown in Fig. 4.23.<br />

Thr^e capacitances are added to the circuit of Fig. 4.20 to obtain the circuit of Fig. 4.23. The<br />

gate-source and gate-drain capacitances, Cgs and Cgd, respectively, represent die junction<br />

capacitance of die gate-to-channel junction. These capacitances are nonlinear — they depend<br />

respectively on gate-drain and drain-source dc voltages, Cgs = J{V G s) an d Qd = J^YGD)- Th e<br />

function J{.) has die form of Equation (3.40) with v D = V GS for Cgs and v D = V DS for Cgd. The<br />

capacitance Cds. is mainly the stray capacitance between the drain and source terminals, so it is a<br />

linear (constant) capacitance to a first approximation.<br />

It should be noted diat small-signal high-frequency model for a MOSFET device is a little more<br />

complicated than the circuit of Fig. 4.23. Namely it comprises diree other capacitances,<br />

connected between the gate, source and drain terminals and die bulk terminal, respectively. These<br />

capacitances are nonlinear (junction) capacitances, unlike die Cgs and Cgd capacitances diat are<br />

stray capacitances for the MOSFET. Such a model will be discussed in Section 4.11 for die<br />

purpose of SPICE simulation of MOSFET circuit behavior. For approximate, simplified analysis<br />

die circuit shown in Fig. 4.23 can be used for both JFET and MOSFET devices.<br />

+ V DS<br />

Figure 4.23 Small-signal high-frequency JFET equivalent circuit<br />

The small-signal JFET equivalent circuits shown in Fig. 4.20 and 4.23 will be used in die next<br />

section to analyze properties of basic types of FET amplifiers at low and high frequencies.<br />

-103-


Analog Electronics / FET Circuits<br />

4.7 Basic Small-Signal FET Amplifier Circuits<br />

The circuit diagram of a common-source amplifier is shown in Fig. 4.24. The ac signal to be<br />

amplified is supplied by %(/). The coupling capacitors C 1 and C 2 as well as the bypass capacitor C s<br />

axe. intended to have very small impedance for the ac signal. In this section we then carry out a<br />

midband analysis of the amplifier, in which we assume that these capacitors are short circuits for<br />

the signal. Later when we consider the frequency response of amplifiers, we include the<br />

capacitors. The resistors R^, R s , and R D form a self-bias network, and their values are selected to<br />

obtain a suitable j2-point. The amplified output signal appears as the voltage drop across the load<br />

resistor R L .<br />

The small-signal equivalent circuit for the amplifier is shown in Fig. 4.25. The input coupling<br />

capacitor has been replaced by a short circuit. The FET has been replaced by its small-signal<br />

equivalent, which is shown in Fig. 4.20. Because the bypass capacitor C s is assumed to be a short<br />

circuit, the source terminal of the FET is connected direcdy to ground - which is why the circuit<br />

is called the common-source amplifier (the source is common to die input and the output). The dc<br />

supply voltage source is considered to be a short circuit for the ac signal. Consequendy, the<br />

resistor R D appears connected from drain to ground.<br />

Figure 4.24 Common-source amplifier<br />

Next we consider the voltage gain of the amplifier. Refer to the small-signal equivalent circuit<br />

and notice that the resistances r d , R D , and R L are in parallel. We denote the equivalent resistance<br />

by<br />

R = - (4.36)<br />

Figure 4.25 Small signal equivalent circuit for the common-source amplifier<br />

The output voltage is the product of the current from the controlled source and the equivalent<br />

resistance.<br />

v 0 = -ig.vJR' L ( 4 ' 3 )<br />

-104-


Analog Electronics /FET Circuits<br />

The minus sign is necessary because of the reference directions selected (i.e. the current g [a v s<br />

flows out of the positive end of the positive end of the voltage reference for pj.<br />

The input voltage and the gate-to-source voltage are equal to each other<br />

Now if we divide Equation (4.37) by (4.38), we find the voltage gain, which is given by<br />

A v =^ = -g m R 1 (4.39)<br />

The minus sign in the expression for the voltage gain shows that the common-source amplifier is<br />

inverting.<br />

The input resistance of the common-source amplifier is given by<br />

R, = — = Rr<br />

(4.40)<br />

h<br />

This impedance forms part of the bias network, but its value is not critical. Practical values<br />

change from 0 to perhaps 10 MQ in discrete component circuits. Thus we have a great deal of<br />

freedom in design of the input resistance of a common-source amplifier. This is not true for BJT<br />

amplifier circuits. One should note, however, that the effective voltage gain<br />

defined as the ratio of the output voltage to the signal source voltage depends on the resistance<br />

jR^. Namely, one can easily demonstrate that<br />

Rr.<br />

K = -g«<br />

R + R, R Rr<br />

L=A V<br />

(4.42)<br />

R + Rr<br />

It follows from Equation (4.42) that with the decreasing gate resistance, the effective voltage gain<br />

decreases, approaching zero when r^—>0. Therefore, high-intemal-resistance sources require<br />

larger values of RQ to maintain large values of the effective voltage gain.<br />

Figure 4.26 Circuit used to find output resistance Ro<br />

To find the output resistance of an amplifier, we disconnect the load, replace the signal source by<br />

its internal resistance, and then find the resistance looking into the output terminals. The<br />

equivalent circuit with these changes is shown in Fig. 4.26.<br />

Because there is no source connected to the input side of the circuit, we conclude that fgs=0.<br />

Therefore, the controlled current source produces 2ero current and appears as an open circuit.<br />

Consequendy, the output resistance is the parallel combination of R D and ra.<br />

R ° = VR D + l/r d ( 4 - 43 )<br />

Exercise 4.8 Find the voltage gain, input resistance and output resistance for the circuit of Fig.<br />

4.24. Also find «(/) and v Q (t). Assume %(/) = 0.1sin(20007t/) V, R = 100 kQ, J^ = 1 MQ, R D = 2.7<br />

-105-


Analog Electronics / FET Circuits<br />

kQ, RL= 10 \sQ, r d = oo, I DSS = 8 mA, V P = -2 V and I D = 2 mA. Ans. A„= -8.5, ft = 1 MQ, Ro =<br />

2.7 kQ.<br />

Exercise 4.9 Find the voltage gain of the amplifier of Exercise 4.8 if an open circuit replaces R^.<br />

Ans. A,,= -10.8.<br />

Exercise 4.10 Find the value of Rs of Exercise 4.8. Ans. R s = 500 Q.<br />

Exercise 4.11 Consider the circuit of Fig. 4.24 with the bypass capacitor replaced by an open<br />

circuit. Draw the small-signal equivalent circuit. Then assuming that ra is an open circuit for<br />

simplicity, derive an expression for the voltage gain in terms of gm and the resistors. Ans.<br />

Exercise 4.12 Evaluate the gain expression found in Exercise 4.11 using the values given in<br />

Exercise 4.8 and the value of Rs found in Exercise 4.10. Compare the results with the voltage<br />

gain found in Exercise 4.8. Ans. A v = -2.84.<br />

Another amplifier circuit known as source follower is shown in Fig. 4.27. The ac signal to be<br />

amplified is supplied by the %(/) signal source and R is the internal resistance of the signal source.<br />

The coupling capacitors C 1 cause the ac input to appear at the gate of the FET. The capacitor C 2<br />

connects the load to the source terminal of the FET. (In the midband analysis of the amplifier, in<br />

which we assume that the coupling capacitors behave as short circuits.) Later when we consider<br />

the frequency response of amplifiers, we include the capacitors. The resistor R s provides a path<br />

for the dc current flowing out of the source terminal of the FET.<br />

Figure 4.27 Source follower<br />

The resistor R^ provides a path for the gate leakage current. One reason for using a source follower is to<br />

obtain high input impedance, and we would pick a large value for R^. The largest resistors available<br />

are on the order of 10 Mfi. Even with such large values, the dc voltage drop caused by leakage<br />

current is usually negligible, so we can consider the bias value of the gate-to-source voltage to be<br />

zero. As a result, the bias value of the drain current is I D = I DSS . Since I DSS demonstrates<br />

considerable device-to-device variation, the bias current of this circuit is not well controlled.<br />

(This situation could be corrected by returning R^ to ground, forming the self-bias circuit<br />

discussed in Section 4.4, but this causes a significant reduction of the input resistance.) Even<br />

though the circuit of Fig. 4.27 has poor bias stability, it achieves extremely high input resistance,<br />

so it is sometimes useful.<br />

The small-signal equivalent circuit is shown in Fig. 4.28. The coupling capacitors have been<br />

replaced by short circuits, and the FET has been replaced by its small-signal equivalent. Notice<br />

-106-


* * = •<br />

(4.44)<br />

Analog Electronics / FET Circuits<br />

that die drain terminal is connected to direcdy to ground because the dc supply becomes a short<br />

in the small-signal equivalent. Therefore, the source follower is also called the common-drain<br />

amplifier. The FET equivalent circuit is drawn in different configuration than shown in Fig. 4.25,<br />

but it is the same electrically.<br />

Drawing the small-signal equivalent for an amplifier circuit is an important skill for electronics<br />

engineers. Test yourself to see if you can obtain the small-signal circuit starting from Fig. 4.27.<br />

Figure 4.28 Small-signal equivalent for the source follower<br />

Now we derive an expression for the voltage gain of the source follower. Notice mat r d , R s and<br />

R L are in parallel. We denote the parallel combination by<br />

1<br />

l/r d +l/R D +UR L<br />

The input current must flow through Rg. Therefore, the current flowing through R L is<br />

''y+SmV-Tkus<br />

v„ = K{L + g^v) (4-45)<br />

We can write the following voltage expression:<br />

V- = V v gS<br />

+<br />

'<br />

V<br />

y l<br />

v 0<br />

Finally,» otice that the voltage across Rg is v^, so we have<br />

v„, = RQif<br />

(4.46)<br />

(4.47)<br />

Equations (4.45), (4.46) and (4.47) form the set needed to solve for the voltage gain. First we use<br />

Equation (4.47) to substitute into Equation (4.45), resulting in<br />

Now if we substitute Equation (4.47) and (4.48) into (4.46), we obtain<br />

v i =R G ii + RL{ii+g m RG i t)<br />

Finally, if we divide Equation (4.48) by (4.49), we find the voltage gain<br />

A = V Q-<br />

R L(l + gm*G)<br />

(4.48)<br />

(4.49)<br />

(4.50)<br />

A few simple checks can be performed on this expression. First, voltage gain is dimensionless.<br />

Checking we see that the expression given on the right-hand side is indeed dimensionless. (Recall<br />

diat the units of g^ are Siemens, i.e. mA/V.) Another simple check is to notice that if ^=0, the<br />

controlled source in Fig. 4.28 becomes an open circuit. Then the equivalent circuit becomes a<br />

-107-


Analog Electronics / FET Circuits<br />

resistive voltage divider. Substituting g^ = 0 into the voltage-gain expression results in the<br />

voltage-division ratio of the resistive circuit. Checks such as this are useful for detecting errors in<br />

writing equations or in the algebra.<br />

Notice that the voltage gain given in Equation (4.50) is positive and is less than unity. However,<br />

in most circuits, it is only slightly less than unity. It follows then, that the output voltage follows<br />

the input voltage that justifies the name source follower for the circuit in Fig. 4.27. To summarize, the<br />

source follower is a noninverting amplifier with a voltage gain slightly lower than unity.<br />

The input resistance can be found from Equation (4.49) by dividing both sides by r,<br />

R i = Vj - = R G^R L {\ + g m R G )<br />

(4.51)<br />

We see that the input resistance can be very large compared to R^.<br />

To find the output resistance, we remove the load resistance, replace the signal source with its<br />

internal resistance, and look into the output terminals. It is helpful to attach a test source v x to the<br />

output terminals as shown in Fig. 4.28. Then the output resistance is found as<br />

R - ^<br />

R "~i<br />

(4.52)<br />

where I x is the current supplied by the test source as shown in the figure. It can be shown that<br />

the output resistance is given by<br />

R„ = ; \ —ir~ (4-53)<br />

— — 1 . + .&A<br />

R + +<br />

s r d R G +R R + R G<br />

This can be quite low, and another reason for using a source follower is to obtain low output impedance.<br />

Figure 4.29 Equivalent circuit used to find die output resistance of<br />

the source follower.<br />

Exercise 4.13 Find the voltage gain, input resistance, output resistance, current gain and power<br />

gain of the source follower shown in Fig. 4.27 if R = 100 kQ, B^ = 10 MQ, R s = 1 kQ, R L = 2.2<br />

kQ. The FET has ra = oo, I DSS =16 mA, V P = -2 V; assume it operates in saturation. Ans. A v =<br />

0.917, ft = 120 MQ, R„= 59.4 Q,A,= 5-10 4 , G = 4.49-101<br />

From the results of Exercise 4.13 we see that even though the voltage gain of the source follower<br />

is less than unity, the output power is much greater than the input power because of the very<br />

high input resistance.<br />

Exercise 4.14 Derive expressions for the voltage gain, input resistance and output resistance of<br />

the source follower shown in Fig. 4.30. Calculate their respective numerical values if R = 100 kQ,<br />

-108-


Analog Electronics / FET Circuits<br />

R G = 10MQ,Rj= 1 kQ.R^ 2.2 kO. The FET has r d =co,I D ss =<br />

operates in saturation.<br />

8 mA, V p = -2 V; assume it<br />

Exetcise 4.15 Derive expressions for the voltage gain, input resistance and output resistance of<br />

the common-gate amplifier shown in Fig. 4.31. Calculate their respective numerical values if R =<br />

100 kQ, Rc= 10 MQ, K s = 1 kfi, Ro= 2.7 kQ, R L = 10 kO. The FET has r d = oo, I DSS = 8 mA, V P<br />

= -2 V; assume it operates in saturation.<br />

TABLE 4.1 PROPERTIES OF BASIC FET AMPLIFIERS<br />

Circuit<br />

configuration<br />

Amplifier<br />

type<br />

Voltage gain<br />

at midband<br />

Input<br />

resistance<br />

Output<br />

resistance Bandwidth<br />

Common<br />

source Inverting -x-R',. Large Medium Medium<br />

Common<br />

drain Noninverting si<br />

Large/,<br />

very large Small Very wide<br />

Common<br />

gate Noninverting<br />

A*«.<br />

Small Medium Wide<br />

Looking at the results obtained from Exercises 4.8, 4.14 and 4.15, one can compare the<br />

properties of the basic FET amplifier configurations. They are summarized in Table 4.1. Please<br />

note that although the common-source and common-gate amplifiers have the same magnitude of<br />

voltage gain, the common-source circuit is an inverting amplifier, unlike the common-gate<br />

configuration. The common-gate one offers larger bandwidth, as will be discussed next in dais<br />

section.<br />

Example 4.1 Find the higher cutoff frequency of the common-source amplifier of Fig. 4.24.<br />

Assume R = 1 kQ, Rc = 10 MQ, R s = 1 kft, R D = 5.6 kfi, R^= 10 kft, r d = oo, I DSS = 16 mA, and<br />

V P = -2 V, Qs = 10 pF, Cgd = 3 pF, Gh = 2 pF.<br />

Solution. To find the circuit small-signal transconductance, one has to determine the j2-point of<br />

the FET by using the techniques of Section 4.4. In the case of the circuit shown in Fig. 4.24, the<br />

gate-to source voltage equals to HGS = -1.407 V. This corresponds to ^ = 2.372 mS. The<br />

equivalent circuit of the amplifier, valid for high frequencies is shown in Fig. 4.32. To build this<br />

-109-


Analog Electronics / FET Circuits<br />

circuit, its model of Fig. 4.23 replaced the FET in Fig. 4.24, and the coupling and bypass<br />

capacitances were replaced by respective short circuits.<br />

Figure 4.32 Small signal high-frequency equivalent circuit for the common-source amplifier<br />

The capacitance Cgd in Fig. 4.32 is connected between the input and the output of the amplifying<br />

device and thus makes the derivations difficult. To simplify the analysis we will use the Miller<br />

theorem of Section 2.6. First, the midband voltage gain should be evaluated between the<br />

terminals to which the feedback capacitance is connected. We will use the circuit of Fig. 4.33 to<br />

find this gain, as<br />

V<br />

A = -- = -g m RL<br />

gs<br />

Since R' L = R D \ \ R L = 3.59 kQ, we find A = -8.52. According to Miller theorem, the capacitance<br />

Cgd may be split into two parts<br />

C m] =C gd (\-A)= 28.55 pF,<br />

\-A<br />

'ml = c gd -A<br />

= 3.35 pF.<br />

These capacitances are connected, respectively, from the gate to source, and drain to source<br />

terminals, as shown in Fig. 4.34. The resistance RQ has been neglected as being much larger than<br />

the signal-source resistance R and the reactance of capacitances (Cgs + C m i) at high frequencies.<br />

Figure 4.34 Simplified equivalent circuit for die common-source amplifier<br />

To find the higher cut-off frequency, one has to evaluate the effective gain of the amplifier<br />

-no-


Analog Electronics / FET Circuits<br />

The output voltage can be described as<br />

^o ~<br />

Sm'gs<br />

l + 7«*/.(C A +C m2 )<br />

(4.54)<br />

R'L<br />

= ~ Sm" gs<br />

1 + JQ)/0)2<br />

where Q) 2 =1/R^C^ +C m2 ) = 1/(3.59 kQ x 5.35 pF) s 1/18.94 ns = 52.8 Mrd/s. The gatesource<br />

voltage is equal to<br />

1<br />

V =V<br />

Y<br />

* gs l + jcoR(C gs +C ml )<br />

=v<br />

1<br />

1 + jco/coi<br />

(4.55)<br />

where


Analog Electronics /FET Circuits<br />

The PSpice simulation run for the circuit of Fig. 4.35a produced^ 2 27.7 MHz. The higher cutoff<br />

frequency of the common-drain amplifier is then much larger than the value of 3.0 MHz<br />

obtained for the common-source amplifier. *<br />

(a)<br />

common-gate amplifier<br />

V 1 0 ac lmV<br />

R 1 2 Ik<br />

Rs 2 0 Ik<br />

Cgs 2 0 lOpF<br />

Cds 2 3 2pF<br />

Cdg 3 0 3pF<br />

gm 3 2 0 2 2.372m<br />

Rd 3 0 5.6k<br />

RL 3 0 10k<br />

.ac dec 100 10Hz lOOMEGHz<br />

.end<br />

Figure 4.36 Small-signal high-frequency equivalent for the common-drain amplifier<br />

Example 4.3 Use PSpice to find the higher cut-off frequency of the common-gate amplifier<br />

shown in Fig. 4.31. Assume R = 1 kQ, R s = 1 kQ, R D ~ 5.6 kQ, R L = 10 kQ. The FET has r d =<br />

oo, I DSS =16 mA, Kp = -2 V, Cgs = 10 pF, Cgd = 3 pF, C


Analog Electronics / FET Circuits<br />

Exercise 4.16. Consider the amplifier shown in Figure 4.37. Both FETs have V& = 3 V, K = 0.6<br />

mA/V 2 , n = oo, Cgs = 10 pF, C g d = 3 pF, Qs = 2 pF, and operate in saturation. Assume R = 1<br />

k£2, R D = 5.6 k£2, R L = 10 kfi. Find device Q-points and transconductance. Draw the small-signal<br />

equivalent circuit valid for medium and high frequencies (replace the capacitances by short<br />

circuits). Run .ac PSpice analysis to find the amplitude characteristic of the amplifier. Determine<br />

the effective gain and bandwidth of the whole cascode as well as gain and bandwidth of its<br />

individual stages.<br />

24V<br />

Figure 4.37 2-MOSFET cascode amplifier<br />

In the next few sections, we turn our attention to other applications for the FET, such as voltagecontrolled<br />

resistance and CMOS logic circuits.<br />

4.8 The FET as a Voltage-Controlled Resistance<br />

Besides its use as an amplifier or as a switch (see Section 4.9), the FET is useful as a voltagecontrolled<br />

resistance. In this application, the bias point is chosen at the origin of the output<br />

(drain) characteristics as illustrated in Fig. 4.38.<br />

10<br />

VDSOO<br />

Figure 4.38 When used as a voltage-controlled resistance, the FET is biased at the origin<br />

If the gate-to source voltage is greater than the pinch-off voltage, the device operates in the linear<br />

(triode) region, and the drain current is given by Equation (4.5), which we repeated here for<br />

convenience<br />

i D =K[2(v GS -V P )v DS -v 2 DS]<br />

The transconductance of the device can be found by application of Equation (4.34), which is<br />

-113-


Analog Electronics / FET Circuits<br />

di r<br />

dv GS<br />

Q-pmnt<br />

Applying this equation to (4.5) we have<br />

°« "»\Q-poml<br />

(4.57)<br />

But the jg-point is v DS = V DS =Q, so we have^n=0. Another way to obtain this result is to recall that<br />

gm is a measure of vertical spacing of the drain characteristics. However, as shown in Fig. 4.38, all<br />

the characteristic curves pass through the origin, and the spacing is zero for that point.<br />

Figure 4.39 Small-signal equivalent circuit for a FET operated at VDS = 0<br />

Because gm = 0, the controlled current source of the small-signal equivalent circuit for the FET<br />

biased as shown in Fig. 4.38 becomes an open circuit. Thus as shown in Fig. 4.39, die smallsignal<br />

equivalent circuit simply becomes a resistor ra connected between the drain and source<br />

terminals. This resistance can be found by application of Equation (4.32), which is repeated here,<br />

for convenience<br />

1 di r<br />

8v DS Q—poinl<br />

Applying this equation to (4.5), we obtain<br />

a<br />

Evaluating for v DS =0 and rearranging, we find that<br />

1<br />

2K(V GS -V P )<br />

(4.58)<br />

(4.59)<br />

which is valid provided that V cs is above pinch-off. Of co. rse, if is less than Vp, die FET is in<br />

cutoff and ra=oo.<br />

Figure 4.40 Voltage-controlled FET attenuator<br />

Thus we see tiiat if the FET is biased at the origin of the drain characteristics, it behaves as a resistor connect<br />

from drain to source, the value of which is controlled by the gate-to-source voltage. This conclusion can also b<br />

made by inspection of Fig. 4.38, where die curves are approximately straight lines intersecting die<br />

origin, and their slopes depend on v GS .<br />

One application of the FET as a variable resistance is in die voltage-controlled attenuator circuit<br />

drawn in Fig. 4.40. The resistor R and die resistance r


Analog Electronics / FET Circuits.<br />

v o= v s<br />

d<br />

R + r d<br />

(4.60)<br />

The control voltage is applied to the gate of the FET. If the control voltage is less than the<br />

pinch-off voltage, ra = oo and no attenuation occurs. However, as the control voltage is raised<br />

above pinch-off, rd becomes smaller and the attenuation becomes greater.<br />

The circuit of Fig. 4.40 is an alternative solution to the semiconductor diode voltage-controlled<br />

attenuator circuit discussed in Section 3.4. The FET-based attenuator is often used to stabilize<br />

die amplitude of RC sine-wave oscillators.<br />

Exercise 4.15 Suppose that R = 10 kQ in the voltage-controlled attenuator in Fig. 4.40 and a<br />

FET having I DSS = 16 mA, V P = -4 V is used. Compute values of ra and A v = *b/i% for a number<br />

of values of v GS = -V c — 0, -1, -2, -3 and -4V. Compare the results with those obtained in Exercise<br />

3.6. ADS. r d = 125 fi, 167 Q, 250 Q, 500 Q and oo ; A v =0.0123, 0.0164, 0.0243, 0.0476 and 1.0,<br />

respectively.<br />

4.9 The CMOS Analog Switch<br />

The CMOS analog switch also known as a transmission gate is shown in Fig. 4.41. We will<br />

see that it acts as a switch that either connects points A and B through a low resistance or<br />

disconnects mem, depending on the digital control signal V c .<br />

We assume that the logic levels for V e are +V DD (high) or -V DD (low). Note mat the control<br />

signal is connected directly to the gate of the NMOS and to the input of the logic inverter. The<br />

inverter output is connected to the gate of the PMOS. The input signal v s to be connected to the<br />

load R L can be either analog or digital and is assumed to range between -V DD and +V DD . The<br />

NMOS substrate is connected to the negative supply. We assume the FETs are identical except<br />

for polarity. The threshold voltage for the NMOS is V^ = V^ and for the PMOS is V^ = -<br />

V A . Furthermore, we assume that K A is less than V D p.<br />

Figure 4.41 CMOS analog switch<br />

If V c = -V DD , both FETs are cut off (provided that v s is within -K DD and + V DI j). Thus for V c<br />

low, an open circuit appears between A and B.<br />

Now consider V c = +K DD . To start, we assume that the input voltage v s is positive, so current<br />

flows from left to right through the FETs. With P S = 0 both the NMOS and PMOS are<br />

conducting. Then as v % increases, current flows through the FETs, raising the output voltage.<br />

-115-


Analog Electronics /FET Circuits<br />

Usually, the resistance of the FETs is low enough so that v Q is approximately equal to v r (The<br />

FETs operate in the triode region.) Notice that the gate-to-source voltage of the NMOS is<br />

V G SN=V C -V 0 (4.61)<br />

However, since we are assuming that V c = V DD , we have<br />

V GSN= V DD-V 0 (4.62)<br />

When the output voltage exceeds V DD -V A , the NMOS becomes cut off, but the PMOS is<br />

heavily conducting. Thus point A is connected to point B by a low-resistance path for all values<br />

of v % between zero and Vjrj D .<br />

Similar reasoning, with source and drain terminals interchanged, applies for negative values of the<br />

input voltage. Thus for V c high, point A is connected to point B regardless of the polarity of v s .<br />

For v 0 between -V DD +V A and V DD -V A , both FETs are on. Outside this range, only one of the<br />

FETs is on.<br />

The resistance of a FET is nonlinear. However, for an analog signal, it is undesirable for die<br />

resistance between points A and B to be nonlinear because this leads to distortion of the output<br />

voltage. Fortunately, it can be shown that the nonlinearities of the PMOS and NMOS cancel in<br />

the range of voltages for which both transistors are on. This is an advantage of the CMOS switch<br />

as compared to circuits having only a single transistor. For an analog input signal taking both<br />

positive and negative values, we should choose positive and negative logic level so that both<br />

transistors are on for the range of input voltages expected.<br />

Exercise 4.16 Suppose that the analog switch of Fig. 4.41 uses enhancement MOSFETs having<br />

\V A \ =1V and IK^lmA/V 2 . Also, V DD =SV and R L is large enough so that v Q is<br />

approximately equal to v s . The control signal is V c = V DDt so die gate is in the on state. Find the<br />

small-signal resistance r d of the NMOS for v s =0, 1, 2, 3, 4, and 5V. Repeat for the PMOS. Find<br />

the effective incremental resistance of each device between points A and B for each input<br />

voltage. (Hint. v DS =0, so Equation (4.56) can be used with appropriate changes in notation.)<br />

4.10 CMOS Logic Circuits<br />

In this section we briefly discuss an important logic family that uses complementary metaloxide-semiconductor<br />

(CMOS) FETs. The term complementary implies that both /(-channel and p-<br />

channel devices are employed.<br />

NMOS inverter with resistive pull-up. Before considering CMOS circuits, we discuss die<br />

simpler inverter circuit shown in Fig. 4.42. The transistor is an »-channel enhancement device<br />

(NMOS) having a threshold voltage V^. The load capacitance represents the input capacitance<br />

of driven circuits (such as gates, inverters).<br />

-116-


Analog Electronics / FET Circuits<br />

The drain characteristics of the NMOS transistor are shown in Fig. 4.43. The input voltage V m is<br />

applied to the gate, so we have v GS = V^. For the moment, we assume that the load capacitance is<br />

an open circuit. Using the values R D =10kQ and V DD =10V we construct the load line shown in<br />

Fig. 4.43.<br />

Notice that if the input voltage is less than the threshold voltage of the transistor (assumed to be<br />

K th =3V in this example), the transistor is cut off. Then the circuit operates at point A, and the<br />

output voltage is K 0 = V DD .<br />

As the input voltage is raised above threshold, the point of operation moves up the load line.<br />

When f^n= V DD , the circuit operates at point B, and the output voltage is low. Thus the circuit<br />

operates as a logic inverter (low input corresponds to high output and vice vend).<br />

Fig. 4.44a shows the transfer characteristic of the MOS inverter with resistive load. Points A and<br />

B are marked, as discussed above. With logic one at the input (v GS = 10 V), the output voltage is<br />

equal to 0.61 V and is considered output logic zero for this inverter circuit At the same time, the<br />

drain current is equal to 0.94 mA. Thus the inverter circuit considered draws a power from a<br />

power supply when its input is logic one. Part of this power is dissipated in die drain resistor [P R<br />

= (10V-0.61V)x0.94mA = 8.8mW] and the rest is dissipated inside the MOS transistor structure<br />

(P T = 0.61Vx0.94mA = 0.6mW). This causes the power loss that shortens the battery life in<br />

portable equipment and leads to heating of the circuit.<br />

-117-


Analog Electronics / FET Circuits<br />

In selecting the value of the pull-up resistor Rp, we encounter conflicting objectives. On die one<br />

hand, we want to make die resistor large because this leads to a small current the transistor is on.<br />

This, in turn, means a smaller demand on the power supply and less heating of die circuit On die<br />

other hand, we want to make R D small, so that when the FET switches off, the load capacitance<br />

is quickly charged. (Usually it is important for logic transitions to take place quickly.)<br />

The CMOS inverter. A solution to diis conflict is to use an enhancement ^-channel MOS<br />

(PMOS) transistor in place of die pull-up resistor as shown in Fig. 4.45. (An additional benefit is<br />

diat die PMOS takes much less chip area dian a resistor and dierefore is advantageous for IC<br />

implementation.)<br />

v DD =\0\<br />

11<br />

In the following discussion we assume diat except for the differences in voltage polarity and<br />

current direction, die NMOS and PMOS have identical characteristics. The direshold of die<br />

NMOS is Vfa n = Vfa which is a positive value, and the threshold voltage for the PMOS is V^ =<br />

-Vfa. Also we assume, as is often the case, that the supply voltage V DD is greater that twice die<br />

direshold voltage magnitude. (In the illustrations, we assume that K th =2V and V DD =\OV.)<br />

Notice in Fig. 4.45 that the source terminal of the PMOS is connected to V DD and that the drain<br />

is connected to the inverter output The gate-to-source voltage of the PMOS is given by<br />

v GS P=^-r DD (4-63)<br />

When V m = V DLh die gate-to-source voltage of the PMOS is zero, so it is cut off. Then it acts as<br />

a very high value of R^, and virtually no current flows from the supply. On the other hand, when<br />

V m = 0, we have P GSP = -V DD , and the PMOS can deliver a large drain current to charge the load<br />

capacitance. Since the NMOS is cut off for UQ^ = V^ = 0, no current flows after die<br />

capacitance is charged.<br />

An important advantage of CMOS logic is that, except during logic transitions, either the NMOS<br />

or the PMOS is cut off, and no current flows. Thus the static power consumption (i.e. the<br />

power consumption of a logic circuit when the logic states are not changing) is virtually zero.<br />

For this reason, CMOS is an attractive choice for battery-operated circuits such as portable<br />

computers.<br />

Now we consider load-line constructions for the CMOS inverter. Recall that in the case of a<br />

resistive load, the load line is straight - the volt-ampere characteristic of a resistor. However, for a<br />

PMOS pull-up transistor, the load line is not straight; instead, it is a characteristic curve of the<br />

PMOS. Furthermore, the resistor results in a fixed load line, but the line (actually, it is a curve)<br />

for the PMOS pull-up changes as V m changes. The load lines are shown for several values of V in<br />

on the characteristics of the NMOS in Fig. 4.46.<br />

-118-


Analog Electronics/FET Circuits<br />

Figure 4.46 Load-line analysis of CMOS inverter.<br />

For the NMOS V^ - 2V and for the PMOS K,hp=-2V.<br />

For V m - 0, die PMOS is highly conductive, but the NMOS if cut off, and the point of<br />

operation is at A. Point B illustrates the operating point for a value of V m greater than the<br />

threshold voltage of the NMOS but less than V DD /2. At point B the NMOS is in saturation and<br />

the PMOS in the triode region.<br />

When V m = V DD /2 = 5V, the load line and the NMOS characteristic intersect not at a single<br />

point but along a line from C to D in Fig. 4.46. Thus as V m increases through V DD /2, the<br />

operating point switches abrupdy from C to D.<br />

Point E illustrates an operating point for a value of V m between V DD /2 and V DD -V A . At point E<br />

die PMOS is in saturation, whereas the NMOS is in the linear region. For V m = K OD , the PMOS<br />

is cut off, and operating point is F.<br />

-119-


Analog Electronics / FET Circuits<br />

K A . The transfer characteristic falls abruptly for V m = V DD /2. The CMOS inverter closely<br />

approximates the ideal transfer characteristic for a logic inverter.<br />

The current flowing from the supply through the transistors of the CMOS inverter, assuming an<br />

ideal open-circuit load is shown in Fig. 4.47b. Notice that if V m = 0 or V m = V DD> then the<br />

current is zero. Maximum current flow occurs for V m = V DD /2. The maximum current value<br />

depends on the supply voltage and on the value of transistors' K.<br />

The CMOS NOR and NAND gates. The circuit diagram of a two-input CMOS NOR gate is<br />

shown in Fig. 4.48a. The source and drain terminals of the FETs are not labeled in the figure.<br />

The devices are physically symmetrical, so either end can be considered to be the source, with the<br />

other becoming the drain. Usually, we consider the source of the PMOS devices to be the<br />

terminal that current enters (i.e., the top en of this circuit). Similarly, we consider the sovirces of<br />

the NMOS devices to be the end that current leaves (Le., the bottom terminals of this circuit).<br />

Designation of source and drain is convenient in analysis of the circuit; however, the physical<br />

construction of a device is the same regardless of which end is the source or drain.<br />

Now we consider the operation of the circuit shown in Fig. 4.48a. If both inputs A and B are<br />

low, the PMOS transistors' M 1 and M 2 are conductive, and both the NMOS transistors M 3 and<br />

M 4 are off. Consequently, the output is high. If either A or B or both are high, one or both of the<br />

PMOS devices are cut off. Furthermore, at least one of the NMOS devices is conductive.<br />

Consequendy, the output is low. Consequendy, this circuit performs the NOR logic function. A<br />

two-input CMOS NAND gate is shown in Fig. 4.48b.<br />

A/ "H MM<br />

Mj<br />

(») (b)<br />

Figure 4.48 CMOS two-input logic gates: (a) NOR gate, (b) NAND gate<br />

Exercise 4.16 For the circuit of 4.48b, prepare a table showing all possible combinations of<br />

inputs (each input can be" high or low), the corresponding state of each transistor, and the<br />

corresponding output. Indicate the state of each transistor either as on for operation in the triode<br />

or saturation region or as o^for operation in cutoff.<br />

Exercise 4.17 Draw the circuit diagram of a three-input CMOS NOR gate.<br />

Exercise 4.18 Prepare a table showing the regions of operation (saturation, triode or cutoff) of<br />

the PMOS and the NMOS for each labeled point on the inverter transfer characteristic shown in<br />

Fig. 4.47.<br />

Exercise 4.19 A CMOS inverter is constructed with symmetrical devices having V^ = 3V and<br />

Vfo. = -3V. Sketch the transfer characteristic to scale if V DD = 15V.<br />

-120-


Analog Electronics/FET Circuits<br />

Exercise 4.20 If the devices have \K\ =lmA/V 2 , find the supply current through the inverter of<br />

Exercise 4.20 if V m = V DD /2. Assume an open-circuit load. Ans. 20.25mA.<br />

4.11 FET Dynamic Circuit Models<br />

Figure 4.49 shows a circuit model for an ^-channel JFET and the corresponding SPICE<br />

parameters. Current source i D describes the static characteristic of the device. In pinch-off, the<br />

drain current of an ^-channel JFET is given by<br />

i D = K(v GS - Vpfil + Xvos) (4.64)<br />

in which K is a proportionality factor that depends on a particular device, V p is the pinch-off<br />

voltage, and the parameter X accounts for the slope of the output characteristic curves in<br />

saturation. In die preceding sections of this chapter we have assumed X = 0 to simplify<br />

discussion. The nonzero value of X leads to a nonzero value of the small-signal output<br />

conductance r


Analog Electronics /FET Circuits<br />

(Additional depletion capacitances C„ and C BD connect source and drain to the substrate in IC<br />

structures.) Ohmic resistances RQ and Ry complete the model. Below there are two examples of<br />

PSpice device statement for JFETs:<br />

Jl Dnode Gnoda Snoda JModName<br />

.MODEL JModNana NJF(VTO>-4 BETA=lE-3)<br />

Jl Dnode Gnoda Snode J2N3819<br />

.LIB EVAL.LIB<br />

Text<br />

notation<br />

TABLE 4.2 SPICE PARAMETERS FOR N-CHANNEL JFKD<br />

SPICE Parameter name Typical Default<br />

notation<br />

value value<br />

v P<br />

VTO Pinch-off voltage -3V -2V<br />

K BETA Transconductance<br />

coefficient<br />

600E-6A/V 2<br />

100E-6 A/V2<br />

X LAMBDA Channel-length<br />

2E-3 V 1 0<br />

modulation coefficient<br />

I IS Saturation current 2E-12 A 1.0E-14A<br />

n N Emission coefficient 1 1<br />

R D<br />

RD Ohmic drain resistance \Cl 0<br />

R s RS Ohmic source resistance 0.1 Q 0<br />

^GDO<br />

CGD Gate-drain depletion<br />

3pF 0<br />

capacitance (zero bias)<br />

^GSO<br />

CGS Gate-source depletion 3.3 pF 0<br />

capacitance (zero bias)<br />

m M Junction grading factor 0.333 0.5<br />

fo PB Built-in barrier potential IV 0.5 V<br />

"Typical values of the parameters are shown for a discrete general-purpose device<br />

An enhancement MOSFET structure is presented in Fig. 4.50. Notice that the length JL and<br />

width W of the channel are labeled on die figure. In the saturation region, the drain current is<br />

given by<br />

i D = j K(v GS - V th ) 2 {\ + Xv DS ) (4.69)<br />

(Previously, we assumed for simplicity that W/L. = 1 and X = 0.) Notice that the current depends<br />

on the width-to-length ratio of the channel. The device designer can vary this ratio to obtain<br />

devices best suited for various functions in a circuit It turns out that the small-signal<br />

transconductance depends on the W/L, ratio as follows<br />

8m = 2^KI D^(\ + XV DS ) (4.70)<br />

whereas the formula (4.65) is valid for MOS as weU. Keep in mind that the equation for ^n and gd<br />

are valid only for operation in the saturation region.<br />

Of course, it is desirable to construct devices with small dimensions so that a large number of<br />

mem fit in a .given chip area. Another advantage of smaller devices is that the device capacitances<br />

are smaller. Provided that the width-to-length ratio is maintained, the current available to charge<br />

and discharge these capacitances is independent of device size. Thus digital circuits constructed<br />

-122-


Analog Electronics / FET Circuits<br />

with smaller MOS transistors can switch faster. Both the speed and complexity of digital MOS<br />

circuits can increase as the device dimensions become smaller.<br />

Figure 4.50 ^-Channel enhancement MOSFET structure with channel length L and channel width W<br />

SPICE uses Equation (4.69) to relate the current to the voltages, except that ohmic resistances<br />

are added in series with the source and the drain. This is shown in Fig. 4. In the MOSFET of Fig.<br />

4.51 there are depletion capacitances C BS and C BD between the substrate and reverse-biased n-<br />

type wells. To describe these capacitances, SPICE uses zero bias capacitances CBS and CBD,<br />

grading coefficient MJ and bulk junction potential PB. The SPICE model also includes the<br />

exponential volt-ampere behavior of the body-drain and body-source junctions. These junctions,<br />

characterized by the reverse saturation current IS, must be kept reverse biased in the simulation (as in the<br />

physical device) by connecting the substrate to the most negative point in the circuit.<br />

Figure 4.51 »-channel MOSFET dynamic circuit model<br />

Fig. 4.51 also suggests capacitance, C GB , between gate and substrate, with the oxide for its<br />

dielectric. It turns out that in places where the gate slighdy overlaps the drain and source<br />

materials there are additional capacitances C G5 and C GD . SPICE computes values for C GS , C GD , and<br />

C GB when we specify the value of oxide thickness, TOX, on the .MODEL line and when W and<br />

L are specified on the element lines of the transistors. TOX = 0.1U, W = 1U, and L = 1U suffice<br />

as rough estimates when actual values are unknown. Resistors R s and R D represent the voltage<br />

drops by current flowing to the external contacts D' and S'. SPICE includes temperature variation in<br />

threshold voltage and pn junction parameters; however, variations in mobility with temperature<br />

are absent in basic versions of SPICE.<br />

-123-


Analog Electronics / FET Circuits<br />

Text<br />

notation<br />

SPICE<br />

notation<br />

TABLE 4.3 SPICE PARAMETERS FOR MOSFETs*)<br />

Parameter name<br />

Typical<br />

Value,<br />

NMOS<br />

Typical<br />

Value, PMOS<br />

Kh VTO Threshold voltage IV -IV<br />

2|K| KP Transconductance<br />

30E-6 A /V 2 12E-6 A/V 2<br />

coefficient<br />

X LAMBDA Channel-length<br />

modulation coefficient<br />

1E-2 V- 1 1E-2 V 1<br />

Ro RD Ohmic drain resistance 10 Q ion<br />

R s<br />

RS Ohmic source resistance 10 Q ion<br />

"Typical values are given for devices having W/L, = 1.<br />

Table 4.3 lists SPICE model parameters for static operation. The capacitances C cs , C CD , and C CB<br />

have been described in the text above. Relationships of the type (4.67) and (4.68) describe the<br />

capacitances Cf B and C DB . Notice that the SPICE transconductance coefficient is<br />

KP = 2|K|<br />

Much work has been expended to obtain SPICE models based on device dimensions and process<br />

parameters. Three of the resulting models are incorporated into PSpice. A particular model can<br />

be selected in the .MODEL statement by the LEVEL=1, 2, or 3 parameter. The default is<br />

LEVEL =1. Detailed discussion of these models is beyond the scope of these lecture notes. The<br />

device statement for MOSFETs takes the form<br />

Mdevice Dnode Gnode Snoda Bnoda MNama L»VALUE W-VALUE<br />

.MODEL MNama NMOS(VTO=l KP=30U LAMBDft-0.01 RD=10 RS-10)<br />

Notice that the first character of a MOS device must be M.<br />

Example 4.4 Write a PSpice code to plot the output characteristics for an NMOS having the<br />

model parameters given in Table 4.3. The device dimensions are JL = 20 um and W — 50 urn.<br />

Assume that the substrate is connected to the source. Allow the drain-to-source voltage to range<br />

from 0 to 20 V in 0.1-V steps and gate-to-source voltage to range from 0 to 10 V in 1-V steps.<br />

Solution. The program listing is<br />

NMOS characteristics example<br />

Ml 2 1 0 0 NAME L=20u W*50u<br />

.MODEL NAME NMOS (VTO=l KP=30u LAMBDA»0.01 RD=10 RS=10)<br />

VGS 1 0 dc 5V '<br />

VDS 2 0 dc 10V<br />

.DC VDS 0 20 0.1 VGS 0 10 1<br />

.END<br />

The reader is suggested to reconstruct the schematic diagram of the simulated circuit from the<br />

code.<br />

After running the program and starting PROBE, we request a plot of ID(M1). The resulting plot<br />

of the output characteristics is shown in Fig. 4.52.<br />

-124-


Analog Electronics /FET Circuits<br />

n IBtmi<br />

Figure 4.52 Output characteristics for Example 4.4<br />

4.12 Summary<br />

Field-effect transistors are 3-state devices that serve as dependent sources in analog applications<br />

and as controlled switches in digital circuits. In one kind of FET, the control gate is insulated<br />

from the conducting channel; another FET class has a gate that is in physical contact with the<br />

channel by electrically separated by a reverse biased or Schottky junction. All ^-channel FETs<br />

share the same state definitions, equations, and circuit models, except for minor notational<br />

differences. Mathematically, the transfer characteristics of all «-channel FETs are the right-hand<br />

branches of parabolas; they differ only in the algebraic sign of the threshold or pinch-off voltage.<br />

Equations for /(-channel FETs are identical to those of //-channel devices; however, output<br />

characteristics plot in the second quadrant instead of the first. The circuit models differ only in<br />

the reference directions of the drain currents. The transfer characteristics are of all />-channel<br />

devices are left branches of parabolas; individual differences are only in the algebraic sign of the<br />

threshold or pinch-off voltage. In analysis of FET circuits we encounter two general classes of<br />

problems: finding ^-points when transistor states are known and finding j2-points when the<br />

states are unknown. The tools and procedures for solving these problems are load lines,<br />

equivalent circuits and guessing and verifying states. The infinite input resistance of the FET<br />

tends to simplify circuit analysis; however, FET circuits usually require us to solve a quadratic<br />

equation and select the solution that has physical meaning. The FET can operate as a voltagecontrolled<br />

linear resistor. The most significant second-order effects in FET are breakdown<br />

associated with reverse-biased junction that places an upper limit on the useful active region,<br />

channel length modulation that causes a nonzero positive slope of output characteristics in the<br />

saturation state, and the decrease of the threshold voltage and channel resistance with<br />

temperature. Internal parasitic capacitances limit the FETs ability to operate at high frequencies.<br />

Some are depletion capacitances associated with reverse-biased junctions; others are linear<br />

capacitances associated with insulated gates. Because the minority charge carriers do not take<br />

significant part in the current conduction in FETs, these devices lack large diffusion capacitances<br />

associated with stored minority charge carriers. SPICE models can simulate the nonlinearities of<br />

the static transistor models, and also include most second-order effects of FETs. Because of high<br />

input impedance and output characteristic that resemble resistor volt-ampere curves in both first<br />

and second quadrants, FETs make excellent switches. Bidirectional transmission gates<br />

constructed from FETs are widely used in both digital and linear applications. For small signals, a<br />

-125-


Analog Electronics / FET Circuits<br />

FET biased in die saturation region functions as a voltage-controlled current source with a<br />

transconductance ^n = 2K(K cr Vth) and the output resistance r 0 = 1/(AJ D ). The common-source<br />

configuration provides a high voltage gain and a very high input impedance, but a limited highfrequency<br />

response. A much wider bandwidth is achieved in the common-gate configuration but<br />

its input impedance is low. The source follower (common-drain amplifier) provides a voltage gain<br />

less than unity but features a low output resistance. Integrated-circuit MOS amplifiers utilize<br />

MOS transistors as amplifying and as load devices. In CMOS technology, both n- and/(-channel<br />

enhancement MOSFETs are used, thus providing the circuit designer with considerable<br />

flexibility. The most important things to remember about FETs are their output and transfer<br />

characteristics, the state definitions, the procedure to find their jg-point given bias circuit, and the<br />

load line concepts, including load lines for CMOS inverter sketched in Fig. 4.46.<br />

Notes:<br />

-126-


Analog Electronics/BJT Circuits<br />

5 Bipolar Transistor Circuits<br />

Bipolar junction transistors (BJTs) are constructed as layers of semiconductor materials (most<br />

often silicon) doped with suitable impurities, producing either »-type or/Mype layer. A simplified<br />

physical structure of an integrated circuit npn transistor is shown in Fig; 5.1a. The actual transistor<br />

action takes place along die A-A' line shown in Fig. 5a. The idealized structure of the bipolar<br />

transistor, corresponding to its cross-section along the A-A 1 line is presented in Fig. 5b, where a<br />

layer of p-type material is placed between two layers of the »-type material. This structure can be<br />

seen as two pn junctions made close together in a single crystal of semiconductor. The current in<br />

one junction affects the current in the other junction.<br />

The layers of the bipolar transistor are called the emitter, the base, and the collector, as shown<br />

in Fig. 5.1b. The circuit symbol of an npn transistor is shown in Fig.5c, including reference<br />

directions for the terminal currents and voltages.<br />

Recall that a pn junction is forward biased with applying positive polarity to the p-side. On the<br />

other hand, reverse bias occurs if the positive polarity is applied to the /|-side.<br />

In normal operation of a BJT as an amplifier, the base-collector junction is reverse-biased and the<br />

base-emitter junction is forward biased. In the following discussion we assume that the junctions<br />

are biased in this fashion unless stated otherwise.<br />

The Shockley equation gives die emitter current i E in terms of the base-to-emitter voltage v BE :<br />

'£ = J ES exp^-l (5.1)<br />

This is exacdy the same equation as for the current in a junction diode given in Equation (3.1),<br />

except for changes in notation. The emission coefficient n is made equal to unity since this is the<br />

appropriate value for most junction transistors. Typical values for the saturation current 1^ range<br />

from 10 12 A to 10" 16 A, depending on the size of the device. Recall that at a temperature of 300K,<br />

the thermal voltage V T is approximately 26mV.<br />

Of course, Kirchhoff s current law requires that the current flowing out of die BJT is equal to the<br />

sum of the currents flowing into it. Thus, referring to Fig. 5.1c, we have<br />

i£ =i C +i B (5.2)<br />

-127-


Analog Electronics /BJT Circuits<br />

This equation is true regardless of the bias conditions of the transistor junctions.<br />

We define the parameter a for the transistor as the ratio of the collector current to the emitter<br />

current.<br />

'C<br />

a = l<br />

(5.3)<br />

E<br />

Values of a range from 0.9 to 0.999, with 0.99 being very typical. Equation (5.3) indicates that<br />

the emitter current is partly supplied through the base terminal and partly through the collector<br />

terminal. However, since a is nearly unity, the collector supplies most of the emitter current.<br />

Substituting Equation (5.1) into (5.3) and rearranging, we have<br />

r V BE<br />

i c = cdgs exp(——)-l (5.4)<br />

For v BE greater than a few tens of a volt, the exponential term in the bracket is much larger than<br />

unity. Then the 1 inside the bracket can be dropped. Also, we define the scale current as<br />

I s =a*ES " (5-5)<br />

and Equation (5.4) becornes (for the normal bias conditions)<br />

V T<br />

(5.6)<br />

Solving Equation (5.3) for i 0 substituting into Equation (5.2), and solving for the base current,<br />

we obtain<br />

i B =(\-a)i E (5.7)<br />

Since a is slighdy less than unity, then only a very small fraction of the emitter current is supplied<br />

by the base. Using Equation (5.1) to substitute for i E , we obtain<br />

i B =(l-a)/ £S exp(^)-l (5.8)<br />

We define the parameter fi as the ratio of the collector current to the base current. Taking the<br />

ratio of Equations (5.4) and (5.8) results in<br />

a<br />

o 'C<br />

(5.9)<br />

P<br />

i B l-«<br />

Values for/ range from about 10 to 1000, and a very common value is /£=100. We can write<br />

-128-


Analog Electronics /BJT Circuits<br />

i c =Pi B (5.10)<br />

Note that since / is usually large compared to unity, the collector current is an amplified version of the<br />

base current. Current flow in an npn BJT is illustrated in Fig. 5.2.<br />

As in the case of the FET, the bipolar transistor is a three-terminal device. There are then three<br />

basic circuit configurations that employ the BJT as an amplifier, i.e. a circuit that has one input<br />

and one output. In each of these configurations, one of the transistor terminals is common to the<br />

input and to the output. The common-emitter configuration for an npn BJT is shown in Fig. 5.3.<br />

The battery connected between the base and the emitter supplies a positive voltage v BE that<br />

forward biases the base-emitter junction. The V& battery produces a positive voltage at the<br />

collector with respect to the emitter. Notice that the voltage across the base-collector junction is<br />

given by<br />

VBC= V BE~ V CE ( 511 )<br />

Thus if VCE is greater than v BE , the base-collector voltage v BC is negative, which is reverse bias.<br />

< *'c<br />

Figure 5.3 Common-emitter circuit configuration for the npn BJT<br />

The common-emitter characteristics of the transistor are plots of the currents i B and i c versus the<br />

voltage v BE and v^. Representative characteristics for a low-power silicon device are shown in<br />

Fig. 5.4.<br />

The common-emitter input characteristic shown in Fig. 5.4a is a plot of i B versus v BE , which<br />

are related by Equation (5.8). Notice that the input characteristic takes the same form as the<br />

forward characteristic of a junction diode. Thus, for appreciable current to flow, the base-toemitter<br />

voltage must be approximately 0.6V. Just as for a junction diode, the base-to-emitter<br />

voltage, for a given current, decreases with temperature by about 2mV/K.<br />

The common-emitter output characteristic shown in Fig. 5.4b is a plot of i c versus v^ for<br />

constant values of i B . The transistor illustrated has /£=100. As long as the collector-base junction<br />

is reverse-biased (ffc^O, or equivalently v C E >t, BE)> we have<br />

; C =/B B =IOOZ B<br />

As VCE becomes less than v BE , the base-collector junction becomes forward-biased, and eventually<br />

the collector current falls as shown at the left-hand edge of the output characteristics.<br />

Refer to Fig. 5.4a and notice that a very small change in the base-to-emitter voltage v BE can result<br />

in an appreciable change in the base current i B , partially if the base-emitter junction is forward<br />

biased, so some current (say, 40|iA) is flowing before the change in v BE is made. Provided that v^<br />

is more than a few tens of a volt, this change in base current causes a much larger change in the<br />

collector current i 0 because i^fii^ In suitable circuits, the change in collector current is<br />

converted into a much larger voltage change than the initial change in v BE . Thus the BJT can<br />

amplify a signal applied to the base-emitter junction.<br />

-129-


Analog Electronics /BJT Circuits<br />

5.1 Load-Line Analysis of a Common-Emitter Amplifier<br />

A simple amplifier circuit is shown in Fig. 5.5. The power-supply voltages V BB and V cc bias the<br />

device at an operating point for which amplification of die input signal ik(t) is possible. We will<br />

show that the amplified version of the input signal appears between the collector and die ground.<br />

~ Figure 5.5 Common-emitter amplifier<br />

The load-line technique will be used to analyze die circuit Applying the Kirchhoff s voltage law<br />

to die base loop, we obtain<br />

V B B + V,(/) = RbhiO + VBEiO (5-12)<br />

-130-


Analog Electronics/BJT Circuits<br />

Figure 5.6 Load-line analysis of the amplifier of Fig. 5.5:<br />

(a) input (load line shifts downwards for a smaller value of«%), (b) output.<br />

A plot of Equation (5.12) is shown in Fig. 5.6a as the load line on the input characteristic of the<br />

transistr -. To establish the load line, we must locate two points. If we assume that i B =0, then<br />

Equatio (5.12) yields v BE =V BB +t>s- This establishes the point where trie load line intersects the<br />

voltage axis. Similarly, assuming that v BE =Q results in i B = (f flB +«fc)/.Rb, which establishes the<br />

load-line intercept on the current axis. The load line is shown in Fig. 5.6a.<br />

Equation (5.12) represents the constraints placed on the values of i B and v BE by the external<br />

circuit. In addition, i B and v BE must fall on the device characteristic. The values that satisfy both<br />

constraints are the values at the intersection of the load line and the device characteristic.<br />

The slope of the load line is -1/Kb. Thus the load line shifts position but maintains a constant<br />

direction ast% changes in value. For example, the lower load line in Fig. 5.6a is for a smaller value<br />

of t>s than that for the upper load line.<br />

The quiescent operating point or Q-point corresponds to %(/)=0. Thus as the ac input signal t%(/)<br />

changes in value with time, the instantaneous operating point swings above and below the Q-<br />

point value. Values of i B can be found from the intersection of the load line with the input<br />

characteristic for each value of t%.<br />

-131-


Analog Electronics /BJT Circuits<br />

After the input circuit has been analyzed to find values of i B , a load-line analysis of the output<br />

circuit is possible. Referring to Fig. 5.5, we can write a voltage equation for the collector loop,<br />

through V cc , R& and the transistor from collector to emitter. Thus we have<br />

YcC = R cic +v CE (513)<br />

This is plotted on the output characteristic of the transistor in Fig. 5.6b.<br />

Now, with the values of i B that we have already found by analysis of the input circuit, we can<br />

locate the intersection of the corresponding output curve with the load line to find values of i c<br />

and «>££. Thus ast% swings through a range of values, i B changes, and the instantaneous operating<br />

point swings up and down the load line on the output characteristic. Usually, the ac component<br />

of V& is much larger than the input voltage; hence the amplification takes place.<br />

Examination of Fig. 5.6a shows that as v&{t) swings positive, the value of i B increases (i.e. the<br />

intersection of the load line with the input characteristic moves upward). This in turn causes the<br />

instantaneous operating point to move upward on the output load line, and v^ decreases in<br />

value. Thus a swing in the positive direction for ik results in a (much larger) swing in the negative<br />

direction for v^. Therefore, as well as being amplified, the signal is inverted. In other words, the<br />

common-emitter amplifier is an inverting amplifier.<br />

>«T : r :.—.,—-^<br />

Figure 5.7 Output of the amplifier of Exercise 5.1 for s (/)=1.2sin(2000iu)<br />

demonstrating gross distortion<br />

Exercise 5.1 Assume that the circuit of Fig. 5.5 has V cc = 10 V, V BB = 1.6 V, R b = 40l£i<br />

R,. = 2 ls£l. The input signal is a 0.4-V peak 1-kHz sinusoid given by t> s (f) = 0.4sin(200(br/). Use<br />

PSpice to plot the device-characteristics assuming fi p = 100 and J s = le-14 A. Using the load-line<br />

technique find the maximum, minimum and Q-point values for v CE . Use PSpice to plot the<br />

waveforms of vjt) and v^t). Ans. p Canki s 3 V, V^ = 5 V, VCmaai s 7 V.<br />

It is not apparent in the waveforms that you have plotted in Exercise 5.1, but the output signal is<br />

not a precise sine wave like the input The amplifier is slightly nonlinear because of the curvature<br />

of the characteristics of the transistor. Therefore, as well as being amplified and inverted, the<br />

signal is distorted. Of course, distortion is not usually desirable. Fig. 5.7 shows the output of the<br />

amplifier of Fig. 5.5 and Exercise 5.1 if the input signal is increased in amplitude to 1.2 V peak.<br />

The distortion is obvious.<br />

and<br />

Notice that the positive peak of v^ has been clipped at K CC =10V. This occurs when i B and i c<br />

have been reduced to zero by the negative peaks of the input signal, and the instantaneous<br />

-132-


Analog Electronics /BJT Circuits<br />

operating point moves down to the voltage-axis intercept of the output load line. When this<br />

happens, we say that the transistor has been driven into cut-off.<br />

The negative-going peak of the output waveform in Fig. 5.7 is dipped at v^ = 0.2 V. This occurs<br />

because i B becomes large enough so that operation is driven into the region at the upper end of<br />

the output load line, where the characteristic curves are crowded-together. We call this the<br />

saturation region.<br />

Reasonably linear (undistorted) amplification occurs only if the signal swing remains in the active<br />

region between saturation and cutoff on the load line. An output load line is shown in Fig. 5.8,<br />

including labels for the cutoff, saturation and active regions.<br />

J IC(Q1)<br />

W2<br />

Figure 5.8 Amplification occurs in the active region. Clipping occurs when the instantaneous operating<br />

point enters saturation or cutoff. In saturation, VCE = 0.2V.<br />

Exercise 5.2 Repeat Exercise 5.1 if Vs{t) = 0.8sin(2(XXbt^. Ans. Pamm = 1 V, V^ = 5 V,<br />

Vcsrv^S.SV.<br />

Exercise 5.3 Repeat Exercise 5.1 if tk(t) = 0.8sin(2000it/) and V BB = 1.2 V. Ans. ^c^un S3V, V^<br />

s 7 V, Pep* s 9.8 V.<br />

5.2 The pnp Bipolar Junction Transistor<br />

So far we have considered the npn BJT, but an equally useful device is the pnp bipolar junction<br />

transistor, in which the base is a layer of »-type material between />-type emitter and collector<br />

layers. For proper operation of an amplifier, the polarities of the dc voltages applied to the pnp<br />

device must be opposite to those of the npn device. Furthermore, current flows in the opposite<br />

direction. Aside from the differences in voltage polarity and current direction, the two types of<br />

devices are nearly identical<br />

A diagram of the structure of zpnp BJT and its circuit symbol are shown in Fig. 5.9. Notice that<br />

the arrow on the emitter of the pnp transistor points into the device, which is the normal<br />

direction of the emitter current For the pnp transistor we can write the following equations,<br />

which are exactly the same as for the npn transistor.<br />

'C = ctiE (5.14)<br />

i B =(\-a)i E (5.15)<br />

ic=fiB (5.16)<br />

-133-


Analog Electronics /BJT Circuits<br />

and<br />

i£ =i C +i B ( 517 )<br />

Equations (5.14) through (5.16) are valid only if the base-emitter junction is forward biased (v BE<br />

negative for i-pnp) and the base-collector junction is reverse biased (y^ positive for a. pup). As for<br />

the npn transistor, typical values are a £ 0.99 and fi= 100.<br />

and<br />

These equations are identical to Equations (5.1) and (5.8) for the npn transistor except that - v#£<br />

has been substituted for v BE (because v BE takes negative values for ihepnp device). As for the npn<br />

device, typical values for 1^ range from 10" 12 to 10" ,6 A, and at 300K we have V-f226tiN.<br />

The common-emitter characteristics ofapnp transistor are exacuy the same as for the npn except<br />

that the values on the voltage axes are negative. A typical set of characteristics is shown in Fig.<br />

5.10.<br />

134


Analog Electronics /BJT Circuits<br />

z.« i<br />

! ItVB)<br />

(b)<br />

Figure 5.10 Common-emitter characteristics for zpttp BJT: (a) input, (b) output<br />

Exercise 5.4 Find the values of alpha and beta for the transistor having the characteristics shown<br />

in Fig. 5.10. Ans. a = 0.98, fi - 50.<br />

Exercise 5.5 Use load-line analysis to find the minimum, maximum and Q-point values of i B and<br />

VCE for the amplifier circuit shown in Fig. 5.11. Use the characteristics shown in Fig. 5.10. Does<br />

this pnp BJT common-emitter amplifier invert the signal Ans. igg^ = 5 |xA, I B = 25 |iA, tg^x =<br />

48 uA, VOa^ £ -8.3 V, KCES -5.3 V, Pcimax s -1.8 V.<br />

1<br />

-12M<br />

i c<br />

Rc=3kCl<br />

5.3 Secondary Effects<br />

Collector breakdown. The description we have given so far is only a first-order model of the<br />

BJT. Real transistors exhibit many secondary effects that can be important in circuit design. For<br />

example, the common-emitter output characteristics of a real transistor are shown in Fig. 5.12.<br />

Notice that the collector current increases very rapidly as the collector-emitter voltage v^<br />

approaches 30V in this case. This is due to the reverse-bias breakdown of the collector-base<br />

junction. Usually, we try to avoid having BJTs enter the collector-breakdown region because high<br />

currents and voltages can result in high power dissipation that leads to overheating and<br />

destruction of the device. Collector breakdown voltages range up to several hundred volts,<br />

depending on the device type.<br />

Base width modulation. Another difference between the first-order BJT model and real<br />

transistors is that even before collector breakdown is reached, the collector current increases with<br />

collector-to-emitter voltage. For example, notice the positive slope of the curves in the active<br />

region of Fig. 5.12a. The slope is more pronounced at higher currents. This effect is attributed to<br />

die base width modulation effect known from the physics of the BJT operation.<br />

-135-


Analog Electronics /BJT Circuits<br />

The shading of Fig. 5.12c and 5.12d suggests how the two depletion regions extend into the/>type<br />

base region during forward active operation. The base-emitter depletion region is narrow,<br />

the reverse-biased base-collector junction has a wider depletion region. For increased v^, the<br />

depletion region of the base-collector junction extends still further into the base, which reduces<br />

the width of the base region. Thus carriers injected from the emitter spend shorter time for die<br />

transit through the base and therefore a lower number of them recombine in the base region.<br />

This brings a closer to one. Since / is a sensitive function of a, the separation of the commonemitter<br />

output characteristics increases noticeably with f^, as illustrated in Fig. 5.12a and 5.12b.<br />

For a constant base current, the collector current increases with v^. This is known as the Early<br />

effect. (The change in common-base output characteristics whose spacing is determined by a, is<br />

hardly noticeable because the percent change in a is small.)<br />

If straight lines extend the collector characteristics in the active region, they (approximately) meet<br />

at a point on the negative v^ axis as shown in Fig. 5.12b. The magnitude of the voltage at the<br />

intersection is called the Early voltage, denoted by V A .<br />

4 «+— '<br />

c mm<br />

si><br />

m<br />

(c)<br />

(d)<br />

Figure 5.12 Common-emitter output characteristic that shows collector breakdown (a);<br />

extensions of die active-region collector characteristics intersect at -VA (b),<br />

base width modulation (c) and (d)<br />

Variation of fi with Q-point. In the first order model of the BJT, the collector characteristic<br />

curves are uniformly distributed in the active region. Real transistors tend to have characteristics<br />

that are crowded closer together at very low and very high currents.<br />

-136-


Analog Electronics /BJT Circuits<br />

At low values of I c , recombination of charge carriers within the depletion region of base-emitter<br />

junction becomes significant. These phenomena manifest themselves by the emission coefficient<br />

n in the Shockley equation for base current having a value close to n = 2 which decreases to n = 1<br />

for medium collector currents (e.g. for /^lmA in Fig.5.13a) whereas the emission coefficient for<br />

the collector current stays constant, n — 1. Thus for low currents, the ratio of base current to<br />

collector current is lower than for medium currents.<br />

At high values of 7 C , the unspoken assumption that the carriers injected into the base and emitter<br />

are not disturbing the original concentration of carriers is no longer valid. These high-injectiondensity<br />

effects, sometimes referred to as "current crowding", cause reduction of p. For large<br />

collector currents, e.g. /^lOmA in Fig. 5.13b, high current density in the base region causes a<br />

significant increase of the minority carrier concentration in the base which, in turn, produces a<br />

masking effect to the built-in nonuniform impurity concentration in -the base. In particular, the<br />

built-in electric field in the base is reduced by the injected charge and carrier transport efficiency<br />

is reduced. This reduces the value of the current gain.<br />

Earlier we defined fi = I C /I B , but since the curves are not uniformly spaced, the value of f) is not<br />

constant for all points in the active region of real transistors. The variation of fi with collector<br />

current is illustrated in Fig. 5.13a for a typical BJT, showing the current gain decrease both at low<br />

and high currents. However, since the curves are generally broad and flat at moderate values of<br />

I a we usually assume that ft is independent of operating point as a first approximation.<br />

-137-


Analog Electronics /BJT Circuits<br />

(c)<br />

(d)<br />

Figure 5.13 Typical variation of ^with collector current (a),<br />

input characteristics of a real transistor (b);<br />

output characteristic at temperature Tt (c), T > T/ (d)<br />

Parameter variation with temperature. Input and output characteristics of the BJT change<br />

with temperature, as do the BJT parameters. The common-emitter input characteristic translates<br />

to the left with increased temperature, in a similar way to the diode forward characteristic shown<br />

in Fig. 3.10. The base current is described by Equation (5.8). As in the diode, V r = kT/q in the<br />

exponent dominates the temperature dependence and causes the voltage V BE to decrease by 2<br />

mV for every degree Kelvin increase in temperature (at constant base current).<br />

The output characteristics increase in separation and translate upward with increasing<br />

temperature as shown in Fig. 5.13c and 5.13d. This reflects the increase of j3 with temperature.<br />

Minority carrier lifetime increases with temperature, increasing the carrier transport efficiency<br />

through the base region and bringing a closer to one. According to Equation (5.9), / is a<br />

sensitive function of a, the result is a large increase in /, as temperature increases. An empirical<br />

relationship that predicts variations in /is<br />

flCn = fiCrJ^\ (5.20)<br />

where T and T R are temperatures in degrees Kelvin and B is a constant called the temperature<br />

exponent. For a class of silicon bipolar transistors one may select B - 1.7, which makes (5.20)<br />

predict that J3 approximately doubles from 27°C to 175°C.<br />

To explain the upward translation, we need to notice that equation (5.10) is an approximation.<br />

Namely, for higher accuracy it should have two terms as follows<br />

where I^g is a small dc saturation current that doubles for every 5K increase in temperature.<br />

From the three temperature-sensitive parameters, V BE actually proves to be the most<br />

troublesome in practical circuits. Even at elevated temperatures, J^ is usually too small to play a<br />

major role in silicon transistors. Also, it is relatively easy to design circuits that work well for any<br />

high value of fi. Many IC designs use the V BE drops of matched transistors to cancel each other<br />

over wide temperature ranges.<br />

Variation of/ from unit to unit. If we test many units of a transistor of a given manufacturer's<br />

type number, we generally "find that J3 displays considerable variation in the value from unit to<br />

unit. Typically, the ratio of the highest and lowest values of beta is 3:1. Furthermore, the value of<br />

beta-varies significantly with temperature for a given transistor. Therefore, m must design cinuits<br />

that junction properly for transistors having a wide range of beta values.<br />

-138-


Analog Electronics /BJT Circuits<br />

Dependence of input characteristics on VCE. The input characteristics of a real transistor are<br />

shown in Fig. 5.13b. Notice that the input charactetistics ate not a single curve, as in the firstorder<br />

model, but instead consist of a family of curves. The effect of base-width modulation with<br />

the varying v K voltage is mainly responsible for this property. However, assuming that v^ is<br />

larger than a few tenths of a volt (i.e. the BJT is biased in the active region), the input<br />

characteristic curves are very close together.<br />

Charge storage effects. The characteristic curves show only die static operation of the device.<br />

For rapidly changing signals, charge storage effects occur, and the instantaneous operation of the<br />

BJT is not adequately described by the characteristic curves. These effects are important in the<br />

design of high-frequency amplifiers and high-speed logic circuits. We consider these aspects later<br />

in this chapter.<br />

Even though real BJTs display many secondary effects that can be important in the design of<br />

critical circuits, the first-order model is sufficient for many designs. Even in critical circuits,<br />

designs often begins with die simple model. In die next few sections, we show some useful<br />

circuits man be analyzed and designed by use of first-order models.<br />

5.4 Large-Signal dc BJT Models<br />

In the analysis or design of BJT amplifier circuits, we often consider the dc operating point<br />

separately from the analysis of the (small) signals. This was illustrated for diode circuits and FET<br />

circuits in previous sections. Usually, we consider the dc operating point first. Then we turn our<br />

attention to the signal to be amplified. In this section, we present models for large-signal dc<br />

analysis of BJT circuits. Then, in die next section, we show how to use these models to design<br />

and analyze bias circuits for BJT amplifiers. Later we consider small-signal models used to<br />

analyze the circuit for the signal being amplified.<br />

It is customary to use uppercase symbols with uppercase subscripts to represent large-signal dc<br />

currents and voltages in transistor circuits. Thus I c and V^ represent the dc collector current and<br />

collector-to-emitter voltage, respectively. Similar notation is used for the other currents and<br />

voltage.<br />

As we have seen, BJTs can operate in the active region, in saturation, or in cutoff. In the active<br />

region, the base-emitter junction is forward-biased and the base-collector junction is reversebiased.<br />

(Actually, the active region includes the forward bias of the collector junction by a few<br />

tenths of a volt.)<br />

C<br />

C<br />

n _ Q<br />

-139-


Analog Electronics /BJT Circuits<br />

Figure 5.14 BJT large-signal models (Note: Values shown are appropriate for<br />

typical small-signal silicon devices at a temperature of 300K.)<br />

Active-region model. Circuit models for BJTs in the active region are shown in Fig. 5.14a. A<br />

current-controlled current source models the dependence of the collector current on the base<br />

current. The constraints given in the figure for I B and K^ must be satisfied to ensure validity of<br />

the active-region model.<br />

0.2V VC£(V) 0.5V vae(V)<br />

(*) (b)<br />

Figure 5.15 Regions of operation on the characteristics of an npn BJT:<br />

(a) output, (b) input characteristic.<br />

Let us relate the active-region model to the device characteristics. Fig. 5.15 shows the<br />

characteristic curves of an npn transistor. The base current I B is positive and p BH =0.7V for<br />

forward bias of the base-to-emitter junction as shown in Fig. 5.15b. Also notice in Fig. 5.15a that<br />

VCE must be greater than about 0.2V to ensure that operation is in the active region (i.e. above<br />

the knees of the characteristic curves).<br />

-140-


Analog Electronics /BJT Circuits<br />

Similarly, for the pnp BJT we must have J B >0 and V CE I(>0.<br />

Cut-off region. In cutoff, both regions are reverse biased, and no current flows in the device.<br />

Thus the model consists of open circuits among all three terminals as shown in Fig. 5.14c.<br />

(Actually, if small forward-bias voltages up to 0.5V are applied, the .currents are non-zero but<br />

often negligible, and we still use the cutoff model.) The constraints on the voltages for the BJT to<br />

be in the cutoff region are shown in the figure.<br />

Inverted mode. When the collector-base junction is reverse-biased and the base-emitter junction<br />

is forward biased, we say that the transistor is operating in the forward or normal mode.<br />

Sometimes, we encounter situations for which the base-collector junction is forward biased and<br />

the base-emitter junction is reverse-biased. This is the opposite of the normal situation, and we<br />

say that the transistor is operating in the inverted mode. Operation in the inverted mode is the<br />

same as in the normal mode, but with the collector and emitter interchanged. Most devices are<br />

not symmetrical, so alpha and beta take different, much lower, values for the inverted mode that<br />

for the normal mode. For now, we concentrate our attention on operation in saturation, cutoff or<br />

the normal active mode regions.<br />

Exercise 5.6 A given npn transistor has /£=100. Determine the region of operation if (a)<br />

7 B =50uA and J^mA, (b) I B =50uA and V CE =5V i (c) V BE =-2V and V CE =-1 V.<br />

Exercise 5.7 A certain/>/#> transistor has /=100. Determine the region of operation if (a) V BE —-<br />

0.2V and K CT =5V, (b) 7 B =50uA and I c =2mA, (c) F CE =5V and 7 B =50uA.<br />

5.5 Large-Signal dc Analysis of BJT Circuits<br />

In this section, we will use the large-signal BJT models presented in Section 5.4 to analyze<br />

circuits.<br />

(a) (b) (c) (d)<br />

Figure 5.16 Bias circuits for Examples 5.1 and 5.2: actual circuit (a),<br />

equivalent circuits assuming operation in cutoff (b),<br />

in saturation (c), in the active region (d)<br />

-141-


Analog Electronics /BJT Circuits<br />

In the dc analysis of BJT circuits, we first assume that the operation of the transistor is in a<br />

particular region (i.e. active, cutoff, or saturation). Then we use the appropriate model for the<br />

device and solve the circuit Next, we check to see if the solution satisfies the constraints for the<br />

region assumed. If so, the analysis is complete. If not, we assume operation in a different region<br />

and repeat until a valid solution is found. (This is very similar to the analysis of diode circuits<br />

using ideal-diode model or a piecewise-linear model.)<br />

This approach is particularly useful in the analysis and design of bias circuits for BJT amplifiers.<br />

The objective of the bias circuit is to place the operating point in the active region so that signal<br />

can be amplified. Because transistors show considerable variation of parameters, such as (3, from<br />

unit to unit and with temperature, it is important for the bias point to be independent of these<br />

variations.<br />

The next several examples illustrate the technique and provide some observations that are useful<br />

in bias-circuit design.<br />

Example 5.1 The dc bias circuit shown in Fig. 5.16a has R B =200kQ, Rc=lkQ, and K CC =15V.<br />

The transistor has /£=100. Solve for I c and VCE-<br />

Solution We will eventually see that the transistor is in the active region, but we start by<br />

assuming that the transistor is cut off (to illustrate how to test the initial guess of operating<br />

region). Since we assume operation in cutoff, the model for the transistor is shown in Fig. 5.14c,<br />

and the equivalent circuit is shown in Fig. 5.16b. We reason that 7 fl =0 and there is no voltage<br />

drop across R B . Hence we conclude that V B£ =15V. However, in cutoff, we must have K fl£ 0 is met, but /3I B >I C is not met.<br />

Therefore, we conclude that the transistor is not in saturation.<br />

Finally, if we assume that the transistor operates in the active region, we use the BJT model of<br />

Fig. 5.14a, and the equivalent circuit is shown in Fig. 5.16d. Solving, we find that<br />

4=(* / ar0.7)/R B =71.5uA<br />

where we have assumed a forward bias of 0.7V for the base-emitter junction. (Some authors<br />

assume 0.6V for low-power silicon devices. Others assume 0.7V.) In reality, the value depends on<br />

the particular transistor and the current level. Usually, the difference is not significant.) Now we<br />

have<br />

lc=Ph = 7.15nA<br />

Finally, V CE =V CC -R < I C = 7.85V<br />

The requirements for the active region are K CE >0.2V and 7 B >0, which are met. Thus the<br />

transistor operates in the active region.<br />

Exercise 5.8 Repeat Example 5.1 with /£=300. Ans. 7 c =14.8mA, K CE =0.2V (saturation).<br />

-142-


Analog Electronics /BJT Circuits<br />

^15V<br />

v CT (V)<br />

(b) P = 300<br />

Figure 5.17 Load lines for Example 5.1 and Exercise 5.8<br />

It is instructive to consider the load-line constructions shown in Fig. 5.17 for the last two<br />

examples. For /£=100, the operating point is approximately in the center of the load line. On the<br />

other hand, for /=300, the operating point has moved up into saturation.<br />

To use this circuit as an amplifier, we would want a Q-point in the active region where changes


Analog Electronics/BJT<br />

Circuits<br />

Exercise 5.11 Solve the circuit shown in Fig. 5.18 to find I c and V^ if (a) >0^5O, (b) /£=150.<br />

Ans. (a) 7c=:0.965mA, ^=-10.35^ (b) 7 c =1.98mA, V a =-0JZV (transistor in saturation).<br />

lOkfl<br />

Figure 5.18 Circuit for Exercise 5.11<br />

In the next example we consider a circuit that achieves an emitter current that is relatively<br />

independent of /3.<br />

W<br />

(b)<br />

Figure 5.19 Circuit of Example 5.2 (a) and its equivalent circuit (b) assuming<br />

operation in the active region.<br />

Example 5.2 Solve for 7 C and V^ in the circuit of Fig. 5.19a if V C( ~=\5V, V BB =5V, Rc=2kQ,<br />

and yS^lOO. Repeat for ^=300.<br />

Solution We assume that the transistor is in the active region and use the equivalent circuit<br />

shown in Fig. 5.19b. Wr»«ing a voltage equation through V BB , the base-emitter junction and R H)<br />

we have<br />

F flB =0.7 + 7 E R H<br />

This can be solved for the emitter current<br />

h= {V BB -0.1)/R E = 2.15mA<br />

Notice that the emitter current does not depend on the value of fi.<br />

Next we can compute the base and collector current using Equations (5.10) and (5.2)<br />

I E = I B +/3I B = (JS+1)I B<br />

Solving for the base current, we obtain<br />

7 B =/ E /09+l)<br />

Substituting values, we obtain the results given in Table 5.1. Notice that 7 B is lower for the higher<br />

P transistor, and I c is nearly constant.<br />

- 144-


Analog Electronics /BJT Circuits<br />

TABLE 5.1 RESULTS FOR EXAMPLE 5.2<br />

1 p JJO^A) J^mA) ^CE(V)<br />

I 100 21.3 2.13 6.44<br />

1 300 7.14 2.14 6.42<br />

Now we can write a voltage equation around the collector loop to find *VCE<br />

Vc^RcIc+Vcz+RzIz<br />

Substituting values found previously, we find that 1^=6.44V for /=100 and<br />

K CE =6.42Vfor^=300.<br />

The Q-point for the circuit of Fig. 5.19a is almost independent of p. However, die circuit is not<br />

usually practical for use in amplifier circuits. First, it requires two voltage sources, V BB and V co<br />

but often one source is readily available. Second, we may want to inject the signal into the base<br />

(through a coupling capacitor), but the base voltage is fixed with respect to ground by the V BB<br />

source. Because the V BB source is constant, it acts as a short-circuit to ground for ac signal<br />

currents (i.e. the V BB source does not allow an ac voltage to appear at the base.)<br />

5.6 Four-Resistor Bias Circuit<br />

A circuit that avoids the objections discussed at the end of the previous section is shown in Fig.<br />

5.20a. We call this the four-resistor BJT bias circuit. The resistors R, and R 2 for a voltage<br />

divider that is intended to provide a nearly constant voltage at the base of the transistor<br />

(independent of transistor p). As we saw in Example 5.2, constant base voltage results in nearly<br />

constant values of I c and K^. Because the base is not directly connected to the supply or ground<br />

in the four-resistor bias circuit, it is possible to couple an ac signal to the base through a coupling<br />

capacitor.<br />

The circuit can be analyzed as follows. First, the circuit is redrawn as shown in Fig. 5.20b. Two<br />

separate voltage supplies are shown as an aid in the analysis to follow, but otherwise the circuits<br />

in part (a) and (b) of the figure are identical Next, we find the Thevenin equivalent for the circuit<br />

to die le t of the dashed line in Fig. 5.20b. The Thevenin resistance R fl is the parallel combination<br />

of R, and R 2 given by<br />

R^1^T<br />

RIM<br />

2 < 5 - 2,)<br />

The Thevenin voltage is<br />

R 2<br />

Vn 'B = V, CC (5.22)<br />

The circuit with the Thevenin replacement is shown in Fig. 5.20c. Finally, its active-region largesignal<br />

model as shown in Fig. 5.20d replaces die transistor.<br />

Now we can write a voltage equation around the base loop of Fig. 5.20d, resulting in<br />

V B = R B I B + V BE + R E I E (5.23)<br />

Of course, for low-power silicon transistors at room temperature, we have K B£ s0.7V. Now we<br />

can substitute<br />

/ £ «=(l+0)/a<br />

and solve to find that<br />

-145-


Analog Electronics /BJT Circuits<br />

VB-VBE<br />

(524)<br />

''-Ri+Q+fiR,<br />

Once I B is known, I E and I c can be easily found. Then we can write a voltage equation around the<br />

collector loop of Fig. 5.20d and solve for V^. This yields<br />

V C E = VCC ~ Wc ~ &Eh (5-25)<br />

« (b)<br />

r cc<br />

07V<br />

'|/ £ =(P+i)4<br />

(c)<br />

(d)<br />

Figure 5.20 Four-resistor bias circuit (a), equivalent circuit show lg separate voltage sources for base and<br />

collector circuits (b), circuit using Thevenin's equivalent for Vcc, R/ and R2 (c),<br />

equivalent to part (c) with active-region transistor model (d).<br />

Exercise 5.12 Find the values of I c and V^ in the circuit of Fig. 5.20a for (a) /fc=100 and (b)<br />

y9=300. Assume that Vg^O.TV and that the circuit elements take on the following values of<br />

R,=10kQ, R,=5kQ, R^lkQ, R E =lkfi, F CC =15V. Ans. (a) J c =4.12mA, V CE =6.12V, (b)<br />

7c=4.24mA, V CE =6.5\V.<br />

Exercise 5.13 Repeat Exercise 5.12 for R,=100kQ and R 2 =50kQ. Compute the ratio of I c for<br />

/=300 to I c for /fc=100 and compare to the ratio of the currents found in Exercise 5.12. Ans.<br />

The ratio of the collector currents is 1.21. On the other hand, in the previous exercise, the ratio<br />

of the collector currents is only 1.029. Larger values of Rl and R2 lead to larger changes in I c<br />

with changes in f3.<br />

We "often use the four-resistor circuit of Fig. 5.20a for biasing BJTs in discrete-component<br />

amplifiers. Now we consider the design of this type of bias circuit The principal problem of bias<br />

circuit design is to achieve nearly identical operating points for the BJTs even though P may vary by a factor of 3<br />

-146-


Analog Electronics / BJT Circuits<br />

more units. Furthermore, some circuits are required to function over a wide range of temperature,<br />

which can cause significant variations in fi and V BE .<br />

Notice that I c and V^ are nearly independent of fi in the circuit of Fig. 5.20a and Exercise 5.12.<br />

This is achieved by selecting values for R, and R 2 that provide a nearly constant voltage to the<br />

base. As the values of R, and R 2 become larger, the Q-point exhibits larger changes with /.<br />

Comparing the results of Exercise 5.13 with those of Exercise 5.12 can see .this.<br />

For the voltage divider to provide a nearly constant base voltage for different values of base<br />

current, the resistors R, and R 2 should be small in value. However, this leads to large currents,<br />

possible overheating, and the need for a larger, more expensive power supply. Moreover, the ac<br />

impedance seen at the base decreases with the decrease of the values of the resistors R, and R 2 ,<br />

which is undesirable in some applications. Thus we also wish to make R, and R 2 as large in value<br />

as possible. As a general rule, a good compromise is to choose R 2 so that the current through it is 10 to 20 times<br />

the largest base current expected.<br />

Equation (5.24) shows that the base current is proportional to the difference between V B and<br />

V BE . Recall that V BE decreases in value by about 2mV/K as temperature increases. Furthermore,<br />

resistor tolerances cause V B to vary. If we design so that the difference between K B and V BE is<br />

very small, these variations could result in troublesome changes in the jg-point. Therefore, we should<br />

design so that V B is much larger than the changes expected in V BE due to temperature variation and the changes<br />

in V B due to resistor tolerances.<br />

Often, we choose V B to be one-third of the supply voltage, which is usually large enough to<br />

ensure a sufficiently stable jg-point. Usually, V B is much larger than V BE , so that the drop across<br />

R E is approximately equal to V B . A rule in common use is to design so that one-third of the<br />

supply voltage is dropped across R^ one-third across the transistor (P^g) and one-third across<br />

R E .<br />

Consideration of the frequency response, peak signal swing, available component values and<br />

various other matters place constraints on the j2-point and the selection of resistor values to be<br />

used in the bias circxiit. Thus the design of the bias circuit is intertwined with the ac performance<br />

of an i iplifier. We consider more aspects of bias design later in conjunction with amplifier<br />

performance.<br />

Figure 5.21 Circuit for Example 5.3<br />

Example 5.3 Suppose that considerations of the frequency response of an amplifier have<br />

dictated that the j2-point of the transistor should be at I c = 2mA and that R^ = 4.7kQ. The<br />

supply voltage is V cc = 25V. The transistor to be used has / ranging from 100 to 300. Design a<br />

four-resistor bias circuit for this amplifier. Use standard 5%-tolerance resistor values.<br />

-147-


Analog Electronics /BJT Circuits<br />

Solution The diagram of the circuit to be designed is shown in Fig. 5.21. The drop across R^ is<br />

V c - I c Rc= 9.4V. This leaves 25 - 9.4 = 15.6V to be allocated between V^ and the drop across<br />

RE-<br />

Suppose that we choose V^ = 10V leaving 5.6V across Rg. Then, since I E = I a we have RE =<br />

V E /I B = 2.8kQ. However, the closest nominal value is R E = 2.7kQ, so that is our choice. (See<br />

Appendix A for nominal 5%-tolerance resistor values.)<br />

Next, we compute the base voltage V B = V BE + V E = 0.7 + 5.6 = 6.3V. The base current is I B =<br />

Z c /p. Since we want the maximum base current to be much less than I 2 (so that V B does not<br />

change excessively when I B changes due to variations in j$), we use the minimum p. Thus I Bmax . —<br />

J c /P = 20|oA. Then using the rule of thumb that I 2 should be 10 times the maximum base<br />

current, we have I 2 = 0.2mA. Now, R 2 = V B /I 2 = (6.3V)/(0.2mA) = 31.5kQ, so we choose R 2 =<br />

33kQ, which is a standard value.<br />

Next, we see that I, = I B + I 2 = 0.22mA and V, = V cc - V B = 18.7V. Finally, R, = V,/I, = 85kQ,<br />

so we choose a close nominal value of R, = 91kQ. (We could just as well have chosen R, =<br />

82k£l)<br />

Thus our design calls for R^ = 2.7kn, R, = 91kfi, R 2 = 33kQ and R^ = 4.7kQ. Of course, many<br />

other choices could have been made in the design, resulting in different but equally useful values.<br />

Usually, there are many rijjfrt answers to the design problems.<br />

Exercise 5.14 Analyze the circuit designed in Example 5.3 to find the jg-point values of I c and<br />

Vcg that actually result with the nominal resistor values for (a) p=100; (b) P=300. Ans. (a) I c =<br />

2.00mA, VCE = 10.1V; (b) I c = 2.13mA, V^ = 9.22V.<br />

Exercise 5.15 In the four-resistor bias network, does I c decrease, increase or stay the same for a<br />

(small) increase in the value of (a) R^, (bJRg, (c) R„ (d) R^, (e) P<br />

Exercise 5.16 In the four-resistor bias network, does K^ decrease, increase or stay the same for<br />

a (small) increase in the value of (a) R& (tyRg, (c) R„ (d) R^, (e) P<br />

Exercise 5.17 Find the maximum and minimum values of I c and V^ for the circuit designed in<br />

Example 5.3. {Hint: Consider the combination of P and values of resistors within ±5% of the<br />

nominal values.) Ans. Icmn. = 2.43mA, i^, = 1.77mA, VcEmax ~<br />

12.1^1^, = 6.76V.<br />

Exercise 5.18 Suppose that V cc = 20V, R^ = lkQ and a j2-point of I c = 5mA is desired. The<br />

transistor has P ranging from 50 to 150. Design a suitable bias circuit. Use standard 5%-tolerance<br />

resistor values.<br />

-148-


Analog Electronics /BJT Circuits<br />

5.7 Small-Signal Equivalent Circuits<br />

Now we turn our attention to small-signal currents and voltages in circuits containing BJTs. First<br />

we establish the notation used in amplifier circuits. We denote total currents and voltages by<br />

lowercase symbols with uppercase subscripts. Thus i B (t) is the total base current as a function of<br />

time.<br />

The Q-point currents and voltages are denoted by uppercase symbols with uppercase subscripts.<br />

Thus I B is the dc base current if the input signal is set to zero.<br />

Finally we denote the' changes in currents and voltages from the Q-point (due to the input signal<br />

being amplified) by lowercase symbols with lowercase subscripts. Thus i t (/) denotes the signal<br />

component of the base current. Since the total base current is the sum of the Q-point value and<br />

die signal component we can write<br />

iB® = h + M (5-26)<br />

Similarly<br />

•taW^re + ^O) ,( 5 - 27 )<br />

The Q-point is established by the bias circuit as discussed in the previous section. Now we<br />

consider how the small signal components are related to each other in the BJT. The total base<br />

current is given in terms of the total base-to-emitter voltage by (5.8), repeated here for<br />

convenience<br />

i B = (1 - ayigs exp(^)-l (5.28)<br />

We are concerned with operation in the active region for which the 1 inside the bracket is<br />

negligible and can be dropped. Substituting (5.26) and (5.27) into the modified (5.28) one obtains<br />

h + ^O = (1" "VES e*P[ VBE t Vbeit) ) (5-29)<br />

V T<br />

This can be written as<br />

I B + i b (l) = (1 - aVus<br />

, ^


Analog Electronics /BJT Circuits<br />

At room temperature, K^=26mV. A typical value of f$ is 100 and a typical bias current for a<br />

small-signal amplifier is l c - 1mA. These values yield r n - 2600Q.<br />

It is easy to show that the signal component of the collector current is given as<br />

» C (0 = A,(0 ' (5.34)<br />

Equations (5.32) and (5.34) relate the small-signal currents and voltages in a BJT. They lead to a<br />

small-signal equivalent circuit of the BJT shown in Fig. 5.22a. This circuit is very useful in the<br />

analysis of the BJT amplifier circuits. It turns out that the pnp transistor has exactly the same<br />

equivalent circuit as the npn - even the reference directions for the signal currents and voltages<br />

are the same. The resistance r„ is given by (5.33) for both transistor types.<br />

An alternative small-signal equivalent circuit is shown in Fig. 5.22b. Instead of the currentcontrolled<br />

current source it uses a voltage-controlled current source to describe the signal<br />

component of the collector current<br />

»c(0 = ft»v te (0 (5.35)<br />

The coefficient g, is called the transconductance of the BJT, defined as<br />

d (<br />

8m<br />

(5.36)<br />

&>BE V T<br />

Q- point<br />

Which one of the circuits shown in Fig. 5.22 is used for circuit analysis is the matter of<br />

convenience.<br />

Figure 5.22 Small-signal equivalent circuit for the BJT<br />

5.8 The Common-Emitter Amplifier<br />

In a BJT amplifier circuit, the power supply biases the transistor at an operating point in the<br />

active region for which amplification can take place. For example, we can use the four-resistor<br />

bias circuit discussed in Section 5.6. Coupling capacitors are used to connect the load and the<br />

signal source without affecting the bias point.<br />

We can analyze amplifier circuits to find gain, input resistance and output resistance by use of the<br />

small-signal equivalent circuit. In this section, and the next we illustrate this procedure for two<br />

important amplifier circuits.<br />

Fig. 5.23a shows the circuit diagram of a common-emitter amplifier. The resistors R,, R 2 , R E<br />

and R^ form the four-resistor biasing network. The capacitor C, couples the signal source to the<br />

base of the transistor, and JC 2 couples the amplified signal at the collector to the load R L . The<br />

capacitor C E is called the bypass capacitor. It provides a low-impedance connection between<br />

the emitter and ground, for ac.<br />

The coupling and bypass capacitors are chosen large enough so that they have very low ac<br />

impedance at the signal frequencies. For simplicity, in our initial small-signal ac analysis, we treat<br />

-150-


Analog Electronics /BJT Circuits<br />

the capacitors as short circuits. However, at sufficiendy low frequencies, the capacitors reduce<br />

the gain of the amplifier, because their impedance increases in magnitude with decreasing signal<br />

frequency.<br />

Vc<br />

V C<br />

(a) Actual circuit<br />

'*><br />

B<br />

(b) Small-signal ac equivalent circuit<br />

'b=0 B<br />

Source<br />

"turned off<br />

^MJ<br />

(c) Equivalent circuit to find Z 0<br />

Figure 5.23 Common-emitter amplifier<br />

Because the bypass capacitor grounds the emitter for ac signals, the emitter terminal is common<br />

to the input source and to the load. This is the origin of the name common-emitter-amplifier.<br />

The analysis we give here is valid for the midband region of frequency. In the low-frequency<br />

region, the effects of the coupling and bypass capacitors must be considered. In the highfrequency<br />

region, a more complex transistor model must be used that includes the frequene<br />

limitations of the transistor. We treat this high-frequency response later in this chapter.<br />

Before we analyze the amplifier, it is very helpful to draw its small-signal equivalent circuit. This<br />

is shown in Fig. 5.23b. The coupling capacitors have been replaced by short circuits and the<br />

transistor has been replaced by its small-signal equivalent.<br />

The dc power supply is replaced by a short circuit. This is appropriate because it has zero internal<br />

resistance, so no ac voltage can appear on it.<br />

Carefully compare the actual circuit of Fig. 5.23a with the small-signal ac equivalent shown in Fig.<br />

5.23b. Notice that the signal source is connected directly to the base terminal because C, has<br />

been treated as a short circuit. Similarly, the emitter is connected direcdy to the ground and the<br />

load is connected to the collector.<br />

-151 -


Analog Electronics / BJT Circuits<br />

Notice that the top terminal of R, connects to the supply in the original circuit, but R, is<br />

connected from base to ground in the equivalent circuit, because the power-supply voltage<br />

source is treated as a short circuit to ground for ac signals. Notice also that R, ends up in parallel<br />

with R^. Similarly, R^ and R L are in parallel. We find it convenient to define R B as the parallel<br />

combination of R, and R 2<br />

Similarly, R^' is a parallel combination of B^ and RL<br />

R'L = - ^ \ - (5-38)<br />

To find the voltage gain v 0 /v/ of the amplifier, first we note that the input voltage is equal to the<br />

voltage across r„, given by<br />

v, = v^ = r K i b (5.39)<br />

The output voltage is produced by the collector current flowing through Rj/<br />

v o =-R L 0i b (5.40)<br />

The minus sign is necessary because of the reference directions for the current and voltage - the<br />

current flows out the positive voltage reference. Dividing (5.40) by (5.39) gives the voltage gain<br />

V Rr<br />

A=-^ = -/*— (5-41)<br />

V- r<br />

v i<br />

'it<br />

Notice that the gain is negative showing that the common-emitter amplifier is inverting. The gain<br />

magnitude can be quite large - several hundred is not unusual.<br />

The expression for gain given in (5.41) is the gain with the load connected. We found the opencircuit<br />

voltage gain useful to characterize amplifiers in Chapter 2. With R L replaced by an open<br />

circuit, the voltage gain becomes<br />

A vo =-/3^ (5.42)<br />

Another important amplifier specification is the input impedance, which in this case can be<br />

obtained by inspection of the equivalent circuit. The input impedance is the impedance seen<br />

looking into the amplifier-input terminals. For the equivalent circuit of Fig. 5.23b, it is a parallel<br />

combination of R B and r n .<br />

Z,=^ = - ^ - (5.43)<br />

', Rli+ r x<br />

In this case the input impedance is a pure resistance. Therefore we can find the impedance by<br />

dividing the instantaneous voltage f, by the instantaneous current i f Of course, if there were<br />

capacitances or inductances in the equivalent circuit, it would be necessary to obtain the<br />

impedance as the ratio of the phasor voltage and the phasor current.<br />

The current gain Aj can be found by use of Equation (2.3). With some changes in notation, the<br />

equation is<br />

4-'-- = 4%- (5.44)<br />

• l i<br />

R L<br />

The power gain G of the amplifier is the product of the current gain and the voltage gain<br />

(assuming that the input and load impedance are pure resistive).<br />

-152-


Analog Electronics /BJT Circuits<br />

G = AjA v (5.45)<br />

The output impedance is the impedance seen looking back from the load terminals with the<br />

source voltage v s set to zero. This situation is shown in Fig. 5.23c. With v s set to zero, there is no<br />

driving source for the base circuit, so i b is zero. Therefore, the controlled source P/ 4 produces<br />

zero current and appears as an open circuit. Thus the impedance seen from the output terminals<br />

is simply R c .<br />

Z 0 = Rc (5.46)<br />

Exercise 5.19 Find A v , A vo , Zn, A\, G, and Z 0 for the amplifier shown in Fig. 5.23a with Rs =<br />

500Q, Ri=10kQ, R 2 =5kQ, R E =lkO, Rc=lkQ, R L =2kft, (3=100, V BE =0.TV, V CC = 15V. If<br />

f,(/)=0.001sin(=-<br />

28.1, G=2980, Z 0 =lkQ, f o =-54.6sin((0t) mV.<br />

Exercise 5.20 Repeat Exercise 5.19 if P = 300. (Hint: do not forget that the Q-point changes<br />

slighdy when beta changes.) Ans. A v =-\09, v4 vo =-164, Z,= 1185Q, Ai=-64.5, G=7030, Z c =lkQ,<br />

v o =-16Jsin((0t) mV.<br />

5.9 The Emitter Follower<br />

The circuit diagram of another type of BJT amplifier called an emitter follower is shown in Fig.<br />

5.24a. The resistors R f , R 2 and R E form the biasing circuit. The collector resistance is not present<br />

in diis circuit. Thus we have a four-resistor biasing circuit with R(- = 0. The input signal is applied<br />

to this circuit through the coupling capacitor C,. The output signal is coupled from the emitter to<br />

the load by the coupling capacitor C 2 .<br />

Vc<br />

V C<br />

(b) Small-signal equivalent circuit<br />

-153-


Analog Electronics / BJT Circuits<br />

(c) Equivalent circuit used to find output impedance Z„.<br />

Figure 5.24 Emitter follower<br />

The ac equivalent circuit is shown in Fig. 5.24b. As before, we replace the capacitors and power<br />

supply with short circuits. The transistor is replaced by its small-signal equivalent.<br />

Notice that as a result, the collector terminal is connected directly to ground in the equivalent<br />

circuit. The transistor equivalent circuit is oriented with the collector at the bottom in Fig. 5.24b,<br />

but it is electrically the same as the transistor equivalent circuit we have used before. Because the<br />

collector is connected direcdy to ground, this circuit is sometimes called a common-collector<br />

amplifier.<br />

The ability to-draw the small-signal equivalents for BJT circuits is an important skill for the<br />

electronic-circuit designer. Carefully compare the small-signal equivalent in Fig. 5.24b to the ori<br />

circuit. Better still; try to draw the small-signal equivalent circuit on your own starting from the<br />

original circuit.<br />

Notice that R, and R 2 are in parallel in the equivalent circuit. We denote the combination by R B .<br />

Also, R E and R L are in parallel and we denote the combination by R L '. In equation form we have<br />

RB ~RI + R,<br />

and<br />

A1A<br />

Rr = RrR E R L<br />

RE+RL<br />

(5.47)<br />

(5.48)<br />

Next we find the voltage gain of the emitter follower. The current flowing through R L ' is 4+Pv<br />

Thus the output voltage is given by<br />

v 0 =R L (\ + {3)i b (5.49)<br />

Writing the voltage equation from the input terminals through r„ and then through the load to<br />

ground, we have<br />

Vi=V b +R L (l + /B)t b (5.50)<br />

Division of (5.49) by (5.50) results in<br />

A- = (5.51)<br />

r K + R L {\ + P)<br />

The voltage gain of the emitter follower is less than unity because the denominator of the<br />

expression is larger than the numerator. However, the voltage gain is usually only slighdy less<br />

dian unity. An amplifier widi voltage gain less than unity can sometimes be useful because it can<br />

have a large current gain.<br />

Also, notice that the voltage gain is positive. In odier words, the emitter follower is noninverting.<br />

Thus if the input voltage changes, the output at the emitter changes by almost the same amount<br />

-154-


Analog Electronics /BJT Circuits<br />

and in the same direction as the input The output voltage follows the input voltage. This is the<br />

reason for the name emitter follower.<br />

The input impedance Zi can be found as the parallel combination of Rg and the input impedance<br />

seen looking into the base of the transistor, which is indicated as Z« in Fig. 5.24b. Thus we can<br />

write<br />

_ RaZu<br />

The input impedance looking into the base can be found by dividing both sides of (5.50) by i b .<br />

Zit=T- = r x +BL L (\ + P) (5.53)<br />

l b<br />

The input impedance of the emitter follower is relatively high compared to other BJT amplifier<br />

configurations. However, if very high input impedance is needed, we often have to resort to<br />

more complex amplifiers using feedback. We consider this approach later. We have already seen<br />

that field-effect transistors are capable of providing much higher input impedance than BJTs.<br />

Once we have found the voltage gain and input impedance of the emitter follower, the current<br />

gain and power gain can be determined by use of (2.3) and (2.5).<br />

The output impedance of an amplifier is the Thevenin impedance seen from the output<br />

terminals. To find the output impedance of the emitter follower, we remove the load resistance,<br />

put the signal source to zero, and look back into the output terminals of the equivalent circuit.<br />

This is shown in Fig. 5.24c. We have attached a test source v x that delivers a current i x to the<br />

impedance we want to find. The output impedance is given by<br />

Zo=T (5-54)<br />

l x<br />

(here again, the impedance can be expressed as the ratio of instantaneous time-varying quantities<br />

because the circuit is purely resistive. Otherwise, we should use phasors.)<br />

To find this ratio, we write equations involving v x and i x . For example, summing current at the<br />

top of J cr, we have<br />

ib+flb+ix-jfe (5-55)<br />

We must eliminate i b from this equation before we can find the desired expression for the output<br />

impedance. We do not want any circuit variables such as i h in the result - only transistor<br />

parameters and resistor values. Thus we need to write another circuit equation.<br />

First, we denote the parallel combination of R„ R, and R 2 as<br />

i<br />

R*RR<br />

R s = „ ' „ (5.56)<br />

The additional equation needed can now be obtained by applying Kirchhoff s voltage law to the<br />

loop consisting of v^ r x and R/.<br />

v x +r„i b +R s i b =0 (5.57)<br />

If we solve Equation (5.57) for t h substitute into (5.55) and rearrange the result, we obtain the<br />

output impedance.<br />

-155-


Analog Electronics /BJT Circuits<br />

v x 1<br />

Z o~ f ~ \+p 1 ( 5 - 58 )<br />

R E<br />

*s+*K<br />

This can be recognized as the parallel combination of R^ and the impedance<br />

z v x R s + r-<br />

«=tr^t<br />

(559><br />

It can be shown that Z ot is the impedance seen looking into the emitter of the transistor, as<br />

indicated in Fig. 5.24c. The output impedance of the emitter follower tends to be smaller than<br />

that of other BJT amplifier configurations.<br />

Exercise 5.21 Compute the voltage gain, input impedance, current gain, power gain and output<br />

impedance of the emitter-follower amplifier shown in Fig. 5.24a. Assume R, = lOkft, R, =<br />

lOOkO, R2=100kQ, R E =2kQ, Ri=lkQ, B=200, K BE =0.7V, V C c=20V. Verify the results using<br />

PSpice. Ans. ^4 V =0.991, Zj=36.5kn, Z 0 =46.6Q, ^4=36.2, G=35.8<br />

In general, the output impedance of the emitter follower is much lower and the input impedance<br />

is much higher than those, of other single-stage BJT amplifiers. Thus we can use an emitter<br />

follower if high input impedance and low output impedance is needed.<br />

Notice that even though the emitter follower gain is less than unity, the current gain is large.<br />

Thus die output power is larger than the input power and the circuit is effective as an amplifier.<br />

If the emitter follower is cascaded with common-emitter stages, amplifiers with many useful<br />

combinations of parameters are possible. Furthermore, there are several other useful amplifier<br />

configurations using the BJT. Later we study additional circuit configurations and consider the<br />

design of multistage amplifiers.<br />

Exercise 5.22. Repeat Exercise 5.21 widi P=300. Compare the results. Verify the results using<br />

PSpice. Ans.y4 v =0.991, Zi=40.1kQ, Zo=33.2Q, ^4=39.7, G=39.4.<br />

5.10 Review of Small-Signal Equivalent-Circuit Analysis<br />

Before we leave the important topic of small-signal equivalent-circuit analysis, we review die<br />

technique and provide a few useful observations.<br />

The first step in analysis is to draw die small-signal equivalent circuit by making the following<br />

changes to the original circuit:<br />

1) Replace the dc power-supply voltage sources by short circuits.<br />

2) Sometimes we may encounter dc current sources. Replace these by open circuits. This is<br />

appropriate because dc current sources force a constant current with no ac component to<br />

flow.<br />

3) If a midband analysis is desired, replace the coupling and bypass capacitors with short circuits.<br />

However, if we want to find expressions for gain or impedance as a function of frequency or<br />

perform a transient analysis, the capacitors should be included in the equivalent circuit (and<br />

we should use phasors to represent currents and voltages in the ac analysis).<br />

4) Sometimes we use inductors to provide a dc connection, but the inductance is picked large<br />

enough so that it has very high impedance for the ac signal. (Usually, this technique is practical<br />

-156-


Analog Electronics /BJT Circuits<br />

only in circuits intended to operate at high frequencies.) Replace such inductors by open<br />

circuits in the small-signal equivalent<br />

5) Replace the transistor with its equivalent circuit If the circuit has several transistors, use<br />

subscripts to distinguish the currents and parameters of different transistors.<br />

It pays to be careful in drawing the equivalent circuit; analysis of an incorrect circuit is time and<br />

effort wasted. Double-check your circuit before writing equations.<br />

Once the small-signal equivalent circuit is finished, we turn our attention to finding expressions<br />

for the gains and impedance of interest First, identify the pertinent currents and voltages and<br />

label them on the equivalent circuit For example, in finding the voltage gain, the pertinent<br />

variables are the input voltage v t and the output voltage tt>. On the other hand, for the input<br />

impedance, we are concerned with v t and the current i f<br />

The output resistance is the Thevenin resistance of the amplifier. To find the output resistance,<br />

we remove the load, turn off independent signal sources and look back into the output terminals<br />

to find the resistance. Turning off the independent signal sources means replacing voltage<br />

sources with short circuits and current sources with open circuits. Dependent sources, such as<br />

me controlled current source of the transistor equivalent, are not turned off - the controlled<br />

source models the effect of the transistor.<br />

Often, it is convenient to attach a test voltage v x to the output terminals as we did in Fig. 5.24c to<br />

find the output resistance of the emitter follower. Then the output resistance is the ratio of v x and<br />

After drawing the small-signal equivalent circuit and identifying the pertinent current and voltage<br />

variables, we use circuit analysis to write equations. Then we use substitutions to eliminate the<br />

unwanted currents and voltages until we have an equation relating the two variables of interest<br />

Make sure that the equations are not dependent. Otherwise, substitution results in cancellation of<br />

all terms, so that you have 0=0 that indicates you must return to writing additional circuit<br />

equations.<br />

After we have found expressions for the gain or impedance, it is a good_idea to check to see that<br />

me units of the expression are correct Voltage or current gain should be. unitless. Input or output<br />

impedance should have units of ohm. In case the units do not check as expected, we should look<br />

for errors in writing the original equations or for algebraic errors.<br />

Small-signal equivalent circuit analysis is not as troublesome as it might seem from this<br />

discussion. We have tried to mention all of the common problems encountered with this<br />

technique so that you will not waste too much time if they come up. Many useful results can be<br />

obtained with ease by the use of small-signal equivalent circuit analysis.<br />

Possibly the thought process and viewpoints gained from the small-signal analysis technique are<br />

just as important as the expressions that we derive using them. After all, if we only wanted the<br />

formulas, we could resort to using a handbook. It is the understanding of the circuits obtained<br />

that makes the technique so important<br />

Example 5.4 A variation of the common-emitter amplifier is shown in Fig. 5.25a. Draw die<br />

small-signal equivalent circuit and derive an expression for the voltage gain.<br />

-157-


Analog Electronics/BJT Circuits<br />

Figure J3.25 Variation of the common-emitter amplifier (a), and<br />

its small-signal equivalent circuit (b)<br />

Solution The small-signal equivalent circuit is shown in Fig. 5.25b. For convenience, we denote<br />

the parallel combination of R^ and R L by R L '. To find voltage gain, we must write equations<br />

involving t>, and v 0 . However, we find it necessary to involve 4, i y and L in the equations. We can<br />

write<br />

v,=W (5-60)<br />

because t> ; is the voltage across r n . Summing currents at the collector node we have<br />

i f =j3i b +i y (5.61)<br />

The output voltage is<br />

v 0 = R L i y (5.62)<br />

Summing voltages around the outside of the circuit yields the fourth equation<br />

Vi = R B i f +v 0 (5.63)<br />

Equations (5.60) through (5.63) are the set we use to find the voltage gain. Before doing algebra,<br />

we check to be sure that enough equations have been written. The variables i h i, and i must be<br />

eliminated. Since a final equation relating t> ( to v 0 must result, a total of four equations is needed,<br />

and that is exactly the number we have written.<br />

Next, we proceed to eliminate the unwanted variables. First, we can solve Equation (5.60) for i b<br />

and substitute into the other equations to obtain the set<br />

If = + h,<br />

' K<br />

V 0 = Rtfy<br />

Vi = R B if+ v o<br />

The first equation in the last set can be used to substitute for i fi resulting in the equation set<br />

v 0 = R L i y<br />

v t =Ri ^ + i 1 + v<br />

\ r * i-ly +V 0<br />

-158-


Analog Electronics /BJT Circuits<br />

Then we solve the first equation in this set for i p substitute into the second equation and use<br />

algebra to form the ratio of v 0 and t>,<br />

v, r„(R L +R B )<br />

Next, we check to make it sure that our expression is unitless, as it should be for the voltage gain.<br />

The check is satisfied. (Recall that P is unidess and r K has the units of resistance.)<br />

Exercise 5.23 Derive the expressions for the input resistance and output resistance for the<br />

circuit of Fig. 5.25a.<br />

where R ot =<br />

R B R S + R B r n + R s r a<br />

(\ + P)R s +r„<br />

Exercise 5.24 The circuit shown in Fig. 5.26 is known as a common-base amplifier. Derive<br />

expressions for the voltage gain, input resistance and the output resistance in terms of P, r x and<br />

die resistor values.<br />

Ans. A y =p-}t)mV,yl^0.819, G=21.<br />

Exercise 5.26. Draw the small-signal equivalent circuits for the circuits shown in Fig. 5.27.<br />

vc<br />

vc<br />

(a) Common-emitter amplifier widi unbypassed emitter resistor<br />

R<br />

Ans. R i =<br />

B +RL<br />

RcRpt<br />

~r,<br />

r RQn+*B<br />

+(l + P)R L " *c + R ot<br />

-159-


Analog Electronics /BJT Circuits<br />

5.11 The Common-Emitter Hybrid-Parameter Small-Signal Model<br />

Another small-signal model for the BJT is shown in Figure 5.28. It is based on a set of two-port<br />

circuit parameters, known as hybrid parameters.<br />

Figure 5.28 Common-emitter A-parameter small-signal equivalent circuit<br />

This equivalent circuit is completely general for small-signal conditions. Given the proper values<br />

of the four parameters, the ^-parameter model accounts for all the second-order effects in the<br />

device. (This is, however, a linear model that does not account for nonlinear effects.) If the<br />

parameters are allowed to be complex-valued functions of frequency, the model is valid for all<br />

frequencies! However, the model parameters are related to the internal device physics in complex<br />

ways, so their variation with frequency is not easy to understand. Consequendy, other models<br />

that are more easily related to the device physics are usually used for high-frequency analysis as<br />

will be seen in the following section.<br />

In tdrms of die ^-parameters, the small-signal currents and voltages are related by<br />

ne =h ie i b + Ke v ce ( 5 - 64 )<br />

i c = h fe i b+ h oe v ce ( 5 - 65 )<br />

-160-


Analog Electronics / BJT Circuits<br />

Notice that h k has the units of resistance, b n and b^ are unitless, and h„ is a conductance.<br />

Starting from Equations (5.64) and (5.65), we can express each parameter as a partial derivative,<br />

evaluated at the operating point. For example, if we set i k — 0 in Equation (5.64) and solve for h^<br />

we have<br />

v be Avj £<br />

h ro =<br />

(5.66)<br />

v ce i b =0<br />

AV CE<br />

Thus we can find the value of h„ by making a small change in v with / fl held constant, and taking<br />

die ratio of the resulting change in P BE and p^. In other words, h„ is the partial derivative of v BE<br />

vidi respect to v^.<br />

Expressions for the other three ^-parameters similar to Equation (5.66) can be found. These<br />

expressions can be used to determine low-frequency values for the A-parameters for the static<br />

characteristics. (This is similar to the procedure we used to determine low-frequency values for<br />

the FET in the previous chapter.) It is important to understand that parameter values found from<br />

the static characteristics of a device are valid only for low-frequency operation. The<br />

characteristics do not account for capacitances.<br />

-OE<br />

Figure 5.29 The A-parameter equivalent circuit with h„ =0 and h K = 0.<br />

The parameter h n accounts for the effect of base-width modulation on the input characteristic of<br />

die device. Typically, its value is very small. Similarly, h x is a small conductance that accounts for<br />

die upward slope of the output characteristics, which is also caused by base-width modulation.<br />

As an approximation, we can set b n and h K to zero. (Since h K is a conductance, setting it to zero<br />

causes it to become an open circuit.) With these changes, the ^-parameter equivalent circuit<br />

reduces to the circuit shown in Figure 5.29. Except for different labeling of the parameters this is<br />

the same as the equivalent circuit of Figure 5.22. Thus we have<br />

K=*K (5-67)<br />

and<br />

hfoSfi (5.68)<br />

We do not intend to make much use of the complete A-parameter circuit in the design or analysis.<br />

However, data sheets sometimes contain information about the ^-parameters. We have included<br />

this discussion primarily so that you will be familiar with thes> parameters when you encounter<br />

them in the literature or on data sheets.<br />

5.12 The Hybrid-* Model<br />

A small-signal equivalent circuit for the BJT known as the hybrid-Tt is shown in Figure 5.30. This<br />

model is motivated by the internal physics of the device. It includes charge storage effects and is<br />

useful over a wide range of frequencies. The resistance r h called die base-spreading resistance,<br />

accounts for the ohmic resistance of the base region. Typically, it is small compared to r x , ranging<br />

from 10 to 100 Q for small-signal devices. Its vale is nearly independent of the operating point.<br />

-161-


Analog Electronics /BJT Circuits<br />

The resistance r K represents the dynamic resistance of the base-emitter junction as seen from the<br />

"internal" base terminal. It is the same as the r K shown in Figure 5.22 and its value is given by<br />

Equation (5.33). The resistance r M accounts for the effects of base-width modulation on the input<br />

characteristic. In other words, r M represents feedback from the collector to the base. In this sense,<br />

it plays virtually the same role that b„ plays in the /^-parameter equivalent circuit The following<br />

approximate formula relates these parameters:<br />

Ke=—— = — ( 5 - 69 )<br />

The value of r M is very large - several megaohms is typical. For simplicity, we often replace it by<br />

an open circuit At high frequencies, this is justified even further because r M is shunted by the<br />

much lower impedance of Cf,.<br />

Figure 5.30 Hybrid-7i equivalent circuit<br />

OE<br />

The resistance r, accounts approximately for the upward slope of the output characteristics of die<br />

transistor. Thus it plays approximately the same role as h m of the ^parameter equivalent circuit.<br />

We can write<br />

r 0 =T- (5.70)<br />

"oe<br />

Sometimes, to simplify analysis, we replace r a by an open circuit<br />

The capacitance C M is the depletion capacitance of the base-collector junction. Its value depends<br />

on the dc base-collector voltage V^ and the device type. Values are often given on data sheets as<br />

Cobo or C Vt . For example, the data sheet for the 2N2222A device lists Cbo value of 8 pF for V^<br />

= -10V.<br />

Sometimes the time constant of the RC circuit between the collector and base terminals is given<br />

on the data sheet For example, the data sheet for the 2N2222A gives the value labeled as r b C f<br />

This time constant is approximately equal to r/7^. Assuming that C M is known, we can use the<br />

value given for the time constant to find r k .<br />

The capacitance C„ accounts for the diffusion base capacitance and junction capacitance of the<br />

base-emitter junction. The value of C K depends on the jg-point and the transistor type. Values<br />

typically range from 10 to 1000 pF for small-signal devices.<br />

Usually, data sheets do:not give values for C K directly. However, the transition frequency^ is<br />

often given. The transition frequency is related to hybrid-7t parameters by the approximate<br />

formula<br />

-162-


Analog Electronics/BJT Circuits<br />

The controlled source gj> x shown in Figure 5.30 accounts for the amplification properties of the<br />

transistor. Using Equation (5.36) we can compute £, from knowledge of the j2-point (and<br />

temperature). It is easy to demonstrate diat for low frequencies and r k = 0, r M = oo and r 0 =


Analog Electronics /BJT Circuits<br />

due to the difference inj2-points. Therefore, we use^ = 300 MHz in computing a value for C^.<br />

Solving Equation (5.71) for C K and substituting values we find that<br />

P 200<br />

C * ~ 2»Sr/r -C„=-<br />

%pF = \96pF<br />

2^x520x300x10'<br />

Finally, the data sheet gives a maximum value for the collector-base time constant of 150 pF.<br />

Thus we have<br />

/fcCy = 150xlO" 12<br />

Solving for r b and substituting the value found for C M , we have<br />

/i=190<br />

Thus we have used the value published in the data sheet to find values for the parameters of the<br />

hybrid-7t equivalent circuit The circuit and values are shown in Figure 5.31.<br />

As you can see, determination of parameter values for a BJT model from the data sheet is not an<br />

exact science. Many parameter, such as E5, show large unit-to-unit variation. Since we must design<br />

circuits that work with all the devices of a given type, an exact model for a particular unit is not<br />

important. Often, we use the worst-case device specification in finding a device model. If our<br />

circuit design meets its goals with a range of device model parameters, including the worst case,<br />

we can be reasonably sure that it can be mass-produced with an acceptable rejection rate.<br />

The hybrid-7i model is useful for die analysis of single- and multistage amplifiers in a wide<br />

frequency range. Slighdy modified, it is incorporated into the PSpice program to represent the<br />

BJT properties for ac analysis.<br />

5.13 Bipolar Transistor Behavior at High Frequencies<br />

The hybrid-7l model can be used to determine the transistor behavior at high frequencies. To<br />

derive the current gain p as a function of complex frequency s, a current source I h (s) is connected<br />

between the base and emitter, and a short circuit is placed between the collector and emitter<br />

terminals, as shown in Figure 5.32. To simplify the analysis, the resistances r u and r, have been<br />

replaced by open circuits.<br />

B<br />

r b B' „ c u C<br />

Figure 5 32 7 quivalent circuit used to derive die short-circuit current gain of die bipolar transistor as a<br />

function of frequency<br />

-164-


Analog Electronics /BJT Circuits<br />

From the K.CL applied to the collector node, we have<br />

lc(s) = gmVA')-sC M V„{s)<br />

i \ (5.74)<br />

For sufficiently low frequencies (still being in the high frequency range) the following inequality<br />

can be fulfilled: &n»sC u , and accordingly Equation (5.74) can be simplified as follows<br />

I c (s) = g m V x (s) (5.75)<br />

On die other hand, from the KVL applied to the input loop one can obtain the relationship<br />

between the base voltage and base current<br />

V x (s) = *„ „ s (5.76)<br />

where &t=1/r& Substituting Equation (5.76) into (5.75) we obtain<br />

i c (s)=— gm !i {s) „ x ( 5 - 77 )<br />

which allows us to derive expression for die short-circuit current gain<br />

Multiplying numerator and denominator of Equation (5.78) by r K and making use of Equation<br />

(5.73) we have<br />

"W-1WC.+C,) «"*><br />

where symbol /, is used to denote low-frequency value of p. Putting s = j


Analog Electronics /BJT Circuits<br />

It follows from Equation (5.82) that the magnitude of the short-circuit common-emitter current<br />

gain P decreases with frequency. This is illustrated by Figure 5.33. At f—fp the gain magnitude is<br />

lower by v 2 (or 3 dB) compared to its low-frequency value p o . This gives an interpretation to<br />

the beta cutoff frequency defined by Equation (5.80).<br />

The parameter^ as described by Equation (5.71) is the common-emitter transition frequency<br />

or unity-gain crossover frequency. In other words, f T is the frequency at which the magnitude<br />

of P is equal to unity. Putting<br />

|A/TO| =<br />

we obtain<br />

A<br />

1+<br />

v 2<br />

\fpj<br />

# = i+ V2 V 2<br />

fp) \fp \jpj<br />

= 1 (5.83)<br />

where the value oifi was assume much bigger than^g. Finally, we have<br />

(5.84)<br />

(5.85)<br />

The value oif T can be found in transistor data sheets. For example, minimum value oif T for the<br />

2N2222A transistor is specified as being equal to 300 MHz. One has to remember that in fact for<br />

a real transistor | P(/^) | is usually larger than unity, due to the current flowing from the base to<br />

collector terminals through capacitance C M , that was neglected under the assumption g m »sC fl .<br />

The parameter^- is thus used to describe the P frequency dependence for frequencies lower than<br />

f T , say foif


Analog Electronics /BJT Circuits<br />

It is easy to find in Fig. 5.23 that the dc base voltage is approximately equal to 1.8 V. Assuming<br />

I/ B£ =0.7 V, we calculate the dc emitter voltage V E - 1 V and the emitter current I E £ 1mA. For a<br />

typical, which is reasonably high, current gain f3 0 we may write<br />

and<br />

V CE =V CC -R C I C -R E I E =5AV<br />

From the transistor data book one can find at I c = 1 mA and V^ = 5 V:<br />

f T = 130 MHz, p 0 = b (e = 600, r Q = 1/A* = 25 kQ. Similarly,<br />

v C B=Vcc-Rck-Vcc-^nr= 4 - 6W -<br />

A] +K2<br />

For this value of V^ one can find from the transistor data book:<br />

C M =C C = 2.6 pF.<br />

The transconductance can be calculated from Equation (5.36) to obtain gn = 38.5 mS. This can<br />

be substituted to Equation (5.85) that can be solved for C„:<br />

Q = 8m •C M s44.5pF.<br />

2nfT<br />

Also from Equation (533) we have r„ S 15.6 kQ. The parameter r h is estimated from the<br />

collector-base time constant as r b = 25 £2.<br />

To simplify the analysis, it is reasonable to apply the Miller theorem to the circuit of Fig. 5.34.<br />

For this purpose we note that A - Vol V„ S -#nR' L - From Equation (5.87), R' L = 4.6 kfi, then A<br />

s -177. Consequendy, C m \ = 2.6x(l+177) s 462.8 pF and Cm2 = 2.6x178/177 = 2.6 pF. Now we<br />

can draw the simplified equivalent circuit for the common-emitter amplifier, where<br />

Ci = C^ + C m] = 507.3 pF, C 2 = C m2 = 2.6 pF. This circuit is shown in Fig. 5.35.<br />

Figure 5.35 Simplified equivalent circuit of the common-emitter amplifier<br />

The schematic diagram of Fig. 5.35 can be further simplified by using Thevenin's theorem to<br />

modify the input circuit, where<br />

r„ + R s +r b<br />

and is equal to 124 Q in the present example, and<br />

v' s =v r n<br />

s<br />

Rs+rb+r*<br />

The resulting final equivalent circuit is shown in Fig. 5.36.<br />

(5.88)<br />

(5.89)<br />

-167-


Analog Electronics /BJT Circuits<br />

For the circuit of Fig. 5.36 we have<br />

Vs<br />

1<br />

V -<br />

1 + JaX'sCi (590)<br />

-y<br />

r n *<br />

**j+'&+'* 1 + 7'®/®!<br />

where ©j = l/(^Cj) s 15.9 Mrd/s, and<br />

y o =-gmV^L7-^~r- (5.91)<br />

where fi>2 = 1/(^1^2) = 83.6 Mrd/s. Combining Equations (5.90) and (5.91) we may write<br />

down the effective voltage gain of the common-emitter amplifier as:<br />

A =YjL = -Po*l> 1 !__ (5.92)<br />

^s<br />

V s R s +1% + r x 1 + joaltox l + ja)/a>2<br />

This is a transfer function of the type discussed in Section 2.6. Using Equation (2.60) and (2.61)<br />

we have<br />

2.12 MHz ,« Q)^ one can write an approximate expression<br />

/ // =® 1 /(2^) = 1/(^C 1 ) (5.94)<br />

Since the gain A in the Miller equations is proportional to the load resistance, the capacitance C,<br />

< •<br />

increases with R^ . Then the upper cut-off frequencyincreases with the Ri decrease. The largest<br />

(<br />

bandwidth is for Ri tending to zero:<br />

On the other hand, it follows from Equation (5.88) that resistance R s depends on the input<br />

signal source resistance. It decreases with R, -• 0. Finally we note that the resistance r K = I c /gn is<br />

t<br />

typically much larger than the base resistance r ¥ We conclude that in the limit of Ri -» 0 and<br />

Rs -> 0, the upper 3-dB frequency of the common-emitter amplifier is equal to the so-called<br />

transverse cutoff frequency<br />

f b = (5.96)<br />

Jb<br />

2nr b (C„+C M )<br />

This frequency determines the largest bandwiddi of a common-emitter amplifier without<br />

feedback. In the present example,_/J =135 MHz.<br />

-168-


Analog Electronics /BJT Circuits<br />

Figure 5.37 Amplitude characteristics of the common-emitter amplifier for different load resistances<br />

With the load resistance decreasing, as the bandwidth increases, the midband gain decreases. This<br />

follows from Equation (5.92) and is illustrated by Fig. 5.37, which is a result of PSpice simulation<br />

performed for the circuit of Fig. 5.35. A summary of PSpice results is presented in Table 5.1.<br />

TABLE 5.1 MIDBAND GAIN AND UPPER 3-DB FREQUENCY<br />

OF THE COMMON-EMITTER AMPLIFIER<br />

Re kQ 5.6 1 0.1<br />

Avso v/v 175 36.7 3.8<br />

fH MHz 2.16 8.14 22.4<br />

teak -. i.e» . i«» iee» ' " v -jJ*f<br />

o ttCttVlCg*)! » wtZZJ/MHiZ) a M«2VMBS3}<br />

.:* frequency<br />

i«o> -<br />

iee»i<br />

Figure 5.38 Magnitude of common-emitter amplifier input impedance as a function of frequency<br />

To find the input impedance dependence on frequency one can use the equivalent circuit in<br />

Fig. 5.35, namely<br />

Zie=r b +:<br />

'x (5.97)<br />

\ + j6)/a)j<br />

where 0)j =\/(r 7r C\).<br />

amplifier is equal to (r^ +r n )<br />

For low frequencies, the input impedance of the common-emitter<br />

that is 15.7 kQ in our example. Zi e decreases with frequency, and<br />

-169


Analog Electronics /BJT Circuits<br />

since the break frequency depends on the Miller capacitance C u it depends also on amplifier load<br />

resistance. For high frequencies, the input impedance goes to r b . This property is sometimes used<br />

to measure the base resistance of bipolar transistors. The PSpice-simulated dependence of the<br />

magnitude of input impedance on frequency, for different load resistances is shown in Fig. 5.38.<br />

Example 5.6 The common-base amplifier of Fig. 5.26 employs a BC548C bipolar transistor. Use<br />

PSpice to compute the effective voltage gain of the amplifier at high frequencies. Assume R, =<br />

100 Q, R, = 68 kQ, R 2 = 12 kQ, RE = 1.1 kQ, Re = 5.6 kQ, R L = oo, Vcc = 12 V.<br />

Solution. Since the^-point of the transistor is the same as calculated in Example 5.5, we can use<br />

the above-obtained values of equivalent circuit elements: r n = 15.6 kQ, ^ = 38.5 mS, C K — 44.5<br />

pF, C M - 2.6 pF, r Q = 25 kQ, r h = 25 Q. The equivalent circuit of this amplifier is shown in Fig.<br />

5.39.<br />

Figure 5.39 Small-signal equivalent circuit of the common-base amplifier<br />

at medium and high frequencies<br />

A PSpice simulation was performed for the circuit of Fig. 5.39. A summary of the simulation<br />

results is presented in Table 5.2.<br />

TABLE 5.2 MIDBAND GAIN AND UPPER 3-DB FREQUENCY<br />

OF THE COMMON-BASE<br />

AMPLIFIER<br />

Re kQ 5.6 1 0.1<br />

41.5 7.7 0.78<br />

fH MHz 9.13 42.4 113.4<br />

Avso v/v<br />

Comparing the contents of Table 5.1 and 5.2, one can notice that the midband gain of the<br />

common-base amplifier is (for the same load resistance) lower than the gain of the commonemitter<br />

circuit. This is due to the fact that we compare the values of the effective voltage gain<br />

A vs =V 0 IV S that takes into account the input signal attenuation in the circuit made of R s and R,,<br />

The input resistance R, for the common-base circuit is much smaller than that of the commonemitter<br />

circuit. Indeed, in the discussed example R, = 31 Q and thus only V* of the signal<br />

amplitude excites the amplifier. On the other hand, the gain A v = V 0 I V i is the same for both<br />

amplifiers. For R^ = 5.6 kQ it is equal to 176 in Examples 5.5 and 5.6. Notice that the commonbase<br />

is a noninverting amplifier, unlike the common-emitter circuit.<br />

The bandwidth of the common-base amplifier is much larger than the bandwidth of the<br />

common-emitter amplifier. This is also seen in' Tables 5.1 and 5.2. The properties of the<br />

-170-


Analog Electronics /BJT Circuits<br />

common-emitter and common-base circuit resemble those of the common-source and commongate<br />

amplifiers, respectively, discussed in Chapter 4. Main differences between them come from<br />

the generally lower transconductance in FETs.<br />

Example 5.7 The common-collector amplifier of Fig. 5.24 employs a BC548C bipolar transistor.<br />

Use PSpice to compute the effective voltage gain and input impedance of the amplifier at high<br />

frequencies. Assume R, = 100 Q, R, = 68 kO, R 2 = 12 kQ, R E = 1.1 kfi, R L = oo, Vcc = 12 V.<br />

Solution. The base-collector voltage V BC is larger in magnitude in this example as compared to<br />

Examples 5.5 and 5.6. To account for this difference, we have C M = 2.1 pF. The other elements<br />

take the previously determined values: r„ = 15.6 kQ, ^ = 38.5 mS, C n — 44.5 pF, r 0 = 25 kQ, r b =<br />

25 Q. The equivalent circuit of this amplifier is shown in Fig. 5.40, where it was assumed that<br />

Rs


Analog Electronics /BJT Circuits<br />

Figure 5.41 Magnitude of the input impedance of the common-collector amplifier<br />

as a function of frequency<br />

One can demonstrate by using analytical derivations applied to the circuit of Fig. 5.40 that its<br />

output impedance is inductive. Then, being very small at low frequencies, it increases in<br />

magnitude with the frequency increases. This behavior is illustrated by Fig. 5.41. Th output<br />

impedance has a small value of about 26 Q at low frequencies; however, its magnitude doubles at<br />

about 50 MHz.<br />

Exercise 5.27. Consider the amplifier shown in Figure 5.43. Both transistors are of the same,<br />

2N2222A, type. Assume Rs = 1 k«, J^ = 1.2 kQ, Re = 1.2 kQ, ^ = 10 kQ. Find device Q-<br />

points and transconductance. Draw the small-signal equivalent circuit valid for medium and high<br />

frequencies. Derive formula for the midband effective gain of this amplifier. Run .ac PSpice<br />

analysis to find the amplitude characteristic of the amplifier using the device model from<br />

EVAL.LIB library. Determine me effective gain and bandwidth of the whole cascode as well as<br />

gain and bandwidth of its Individual stages.<br />

-172-


Analog Electronics/BJT Circuits<br />

5.14 Large-Signal Dynamic Model for the Bipolar Transistor<br />

In Section 5.4 we presented simple large-signal dc models for the BJT in the various operating<br />

regions. These models are shown in Figure 5.14 and are appropriate for analysis of dc circuits. A<br />

nonlinear large-signal model for the BJT, known as the Ebera-Moll model is shown in Figure<br />

5.44. This model is suitable for computer simulation of the transistor static behavior in all the<br />

regions of operation. The diodes D E and D c model the base-emitter and base-collector junction,<br />

respectively. In parallel with each diode is a current-controlled current source, and the controlling<br />

current is the current in the other diode. This model is valid for low-frequency signals in all four<br />

operating regions.<br />

BO<br />

apiDE<br />

ORiDC<br />

Figure 5.44 Ebers-Moll static model<br />

We have used subscripts to distinguish a F and a R . Previously, we have been concerned mainly<br />

with the forward active region, and we used the symbol a instead of a P In equation form<br />

c<br />

forward active region<br />

In a similar fashion, we have<br />

l E<br />

'C<br />

reverse active region<br />

(5.98)<br />

(5.99)<br />

-173-


Analog Electronics /BJT Circuits<br />

We have used the = sign in Equations (5.98) and (5.99) because these expressions neglect the<br />

reverse leakage currents of D B and D 0<br />

Equation (5.9) gives (3 in terms of a. We can write similar equations for fipin terms of a F and for<br />

/ R in terms of a R . Thus we have<br />

P F =-^— (5.100)<br />

\-a F<br />

and<br />

A-r£" ( 5101 )<br />

\-a R<br />

Typical values for this parameters are a = a F = 0.99, 0 = fip - 100, a R = 0.5, fi R = 1. Let us<br />

consider die physical basis for dieses parameter values, assuming an npn device. Transistors are<br />

usually designed to be operated in die forward active region. Thus the doping of die emitter is<br />

much heavier dian that of die base. Furthermore, the collector junction has a relatively large area<br />

to collect the minority carriers (electrons) diffusing dirough die base. Moreover, due to<br />

nonuniform base doping^ there is an electric field "built-in" in die base which accelerates<br />

electrons toward die collector, reduces their transit time dirough me base and dius decreases die<br />

electron loss due recombination. These features ensure that almost all of the carriers crossing die<br />

emitter-base junction are swept into the collector. Thus in the forward active region, i c = i E , and<br />

a F is close to unity.<br />

In die reverse active region, die base-collector junction is forward biased and die base-emitter<br />

junction is reverse-biased. The collector is not as heavily doped as die emitter is. Thus a<br />

significant fraction of the current crossing die collector-base junction is due to holes crossing<br />

from me base into die collector. Furthermore, the electrons mat are injected into die base region<br />

take longer to diffuse to die emitter junction. The emitter junction is smaller and it takes longer<br />

for die electrons to "find" the junction and be swept into the emitter. Consequently, more of die<br />

electrons recombine in die base. This is why i E is typically only half of i c in die reverse active<br />

region.<br />

The currents in diodes D c and D E of die Ebers-Moll model are given by the Schockley equation.<br />

Thus we have<br />

4tY<br />

»D£=/£Sexp|^-)-l (5.102)<br />

and<br />

'DC = /csexp^^|-l (5.103)<br />

The collector current and emitter current are given by<br />

i c = a F i DE -ipc (5.104)<br />

and<br />

'£ = -


Analog Electronics/BJT Circuits<br />

where 7, is known as the device scale current, firstintroduced in Equation (5.5). Since a F = 1, we can<br />

write<br />

I s = IES ( 5 - 108 )<br />

Therefore, these symbols are sometimes used interchangeably.<br />

Figure 5.45 Dynamic BJT model<br />

Base and Collector Resistance. All components in an integrated circuit are fabricated on a<br />

single piece of semiconductor material called the substrate. A consequence is that external<br />

connections to the base, emitter and collector are located at the top of the IC as shown in Fig.<br />

5.1a. The transistor in this figureis a vertical transistor because current flows vertically through the<br />

active region beneath the emitter material. Three parasitic resistances are thus important in the<br />

planar bipolar transistor. The first,the base spreading resistance r h is the ohmic resistance of the base<br />

current path from the junction to the base contact at the surface. This parameter has been<br />

accounted for in the hybrid-n model of the transistor. It is important in high-current transistors<br />

and significantly influences amplifier frequency response. Next in importance is the ohmic<br />

resistance of the collector, r c . Because N D in the collector is relatively low to reduce base width<br />

modulation, the collector region has rather high resistivity. The result is a collector resistance of<br />

the order 10 - 100 Q that may cause internal transistor saturation due to the i c r c voltage drop on<br />

it A special highly conduction buried layer helps reduce r c . Of lesser importance than the other<br />

two is r e , the ohmic resistance in the highly doped emitter material. A typical value is 1 CI. Figure<br />

5.45 shows a large-signal transistor model, which comprises the Ebers-Moll model with parasitic<br />

resistances added.<br />

Parasitic Capacitances. We know that diodes possess parasitic internal capacitances, nonlinear<br />

Q versus V relationships that are transparent to slow signals but important when rapid changes in<br />

voltage or current are imposed upon the device. Since the BJT contains two pn junctions, we<br />

expect it to have similar dynamic limitations. Associated with each junction are depletion and<br />

diffusion capacitances, which limit high frequency performance. In forward active region,<br />

depletion capacitance is dominant at the reverse-bias collector-base junction. At the forwardbiased<br />

base-emitter junction, diffusion capacitance and depletion capacitances are both<br />

important<br />

-175-


Analog Electronics /BJT Circuits<br />

The transistor depletion capacitances are exactly those we described in diode theory. Thus the<br />

parameter values in Equation (3.40) and the capacitance characteristic within the transistor<br />

depend upon the individual junction grading and geometry.<br />

The diffusion capacitance for a transistor differs slightly from that of an isolated diode because of<br />

the narrow base. For the forward active transistor, the minority charge concentration profile in<br />

the base is triangular as in Figure 5.46. This is because base width W is much smaller than the<br />

electron diffusion length, and because the electric field at the collector junction sweeps away<br />

electrons as fast as they arrive, pinning down the concentration at the collector end of the base to<br />

zero.<br />

Emitter 0 Base W Collector<br />

Figure 5.46 Excess minority charge stored in the base of a forward active npn transistor<br />

For a base-emitter junction of cross-sectional area A, we use Figure 5.46 to calculate that the<br />

quantity of minority charge stored in the base is<br />

Q FA =qA~n[Q)W (5.109)<br />

Diode theory tells us that the electron concentration at the emitter end of the base is controlled<br />

by the base-emitter junction voltage v BE<br />

where N cb is the concentrations of acceptor ions in the base material. Substituting Equation<br />

(5.110) into (5.109) gives the charge-voltage relationship that characterizes the diffusion<br />

capacitance of the forward-active transistor<br />

QFA =<br />

qAW nf<br />

-fe)<br />

(5.111)<br />

2 N ab yr T<br />

This particular collection of stored charge represents electrons in transit from emitter to<br />

collector. On the average, these injected electrons take T F seconds to traverse the base, where T F<br />

is called the forward transit time. For integrated npn transistors, the transit time is of the order<br />

of 1 ns, for integrated/>»p transistors T F = 30 ns. It can be shown that the forward transit time is<br />

related to the transition-frequency by<br />

*'-W T (5 - 1,2 ><br />

Similarly to the forward transit time, r R is the average time that a minority carrier spends in the<br />

base region for operation in the reverse active region. It can be shown that<br />

P F r F =P R r R (5.113)<br />

Figure 5.45 introduces a dynamic BJT model. We recognize it as the Ebers-Moll model of Fig. 5.44<br />

augmented by nonlinear diffusion (subscript d) and depletion (subscript J) capacitances. Each<br />

diffusion capacitance is a nonlinear Q versus V relation, which for the base-emitter diffusion<br />

capacitance takes the form of Equation (5.111). In amplifiers, these internal capacitances reduce<br />

-176-


Analog Electronics /BJT Circuits<br />

the high-frequency gain. In digital circuits, the capacitances introduce switching delays that will<br />

be considered in the following section.<br />

SPICE Parameters for the BJT. Table 5.3 lists basic parameters required to model the static<br />

and dynamic behavior of the transistor using PSpice. It includes second-order effects, such as the<br />

Early effect and ohmic resistances. As shown in the table, the parameters fall into three groups.<br />

First are the parameters that model the static characteristics of the device (IS, BF, BR, RB, RC,<br />

RE, VAF). The second group of parameters models the depletion capacitances of the two<br />

junctions (CJC, MJC, VJC, CJE, MJE, VJE). These parameters are similar to those discussed in<br />

Section 3 for/w-junction diodes. The third group of parameters (TF, TR) models charge storage<br />

in the base region.<br />

Text<br />

notation<br />

Tal Die 5.3 SPICE Model Parameters for the BJT*)<br />

SPICE Parameter name Typical<br />

notation<br />

value<br />

Default<br />

value<br />

I, IS Scale current 1E-14A 1.0E-16A<br />

A<br />

BF Forward beta 100 100<br />

A<br />

BR Reverse beta 1 1<br />

h RB Ohmic base resistance 10 Q 0<br />

r c RC Ohmic collector resistance \Cl 0<br />

r, RE Ohmic emitter resistance 0.1 Q. 0<br />

v A<br />

VAF Forward Early voltage 100 V QO<br />

CJC B-C depletion capacitance 10 pF 0<br />

TF Forward transit time 500 ps _ 0<br />

„. \- ,._<br />

TR Reverse transit time 50 ns J 0<br />

*Typical values of the parameters are shown for a discrete general-purpose device<br />

5.15 Switching Behavior of the Bipolar Junction Transistor<br />

Many transistor applications such as digital logic gates, interfacing circuits, power supplies and<br />

communication circuits use the BJT as a switch operated by a control signal. In such applications<br />

the transistor operates as a two-state device, with saturation corresponding to a closed switch.<br />

In this section we discuss briefly the switching behavior of the BJT. The primary objective of this<br />

section is to relate switching behavior to the internal device physics. Consider the simple RTL<br />

inverter shown in Figure 5.47. We use the PSpice code as listed in the Figure to analyze the<br />

circuit and generate the waveforms shown in Figure 5.48.<br />

-177-


Analog Electronics /BJT<br />

Circuits<br />

^cc= + 3V<br />

BJT switching bahaviour<br />

Vin 1 0 PWL(0 0 O.lu O O.llu 3 0.3u 3 0.31u O)<br />

KB 1 2 5K<br />

Ql 3 2 0 Q2N2222A<br />

.LIB EVAL.LIB<br />

RC 4 3 2k<br />

Vec 4 03V<br />

.TBAN lOOn lu 0 In<br />

.and<br />

Figure 5.47 RTL inverter<br />

The input voltage v iH is the 3-V pulse shown in Figure 5.48a. Initially, the input voltage is zero,<br />

and the transistor is in thecutoff region. Therefore, the base current is 2ero, the collector current<br />

is zero and the output voltage v t is 3 V.<br />

During the time interval from / = 0.10 )j.s to / = 0.11 |4.s, v- m rises rapidly to 3 V. The immediate<br />

effect is to cause i B to increase rapidly. This is shown in Figure 5.48b. The current that flows into<br />

-178-


Analog Electronics /BIT Circuits<br />

the base charges die B-E junction depletion capacitance and dius raises the base-to-emitter<br />

voltage.<br />

Part of the base current flows through the collector junction capadtance and out die collector<br />

lead (opposite to the usual current direction for an npn transistor in die active region, see Figure<br />

5.38b). This current causes die output voltage to increase. Notice diat die output voltage actually<br />

goes slighdy higher dian die supply voltage (which is 3 V).<br />

Shortly after the beginning of die input pulse, tine base voltage rises high enough to forward bias<br />

die emitter junction. Then electrons cross from the emitter into the base. These electrons diffuse<br />

into me collector junction. Thus conventional current begins to flow into the collector, and die<br />

collector voltage v, starts to drop. At about / = 0.19 ps, die transistor enters the saturation region<br />

(P's > 'o c f- Figure 5.48b). Then die collector voltage becomes approximately constant at a few<br />

tendi of a volt.<br />

The input switches back to zero during the time interval from / = 0.30 u.s to / = 0.31 us.<br />

However, die output voltage remains low until approximately / = 0.52 u.s. The reason for diat is<br />

die excess minority carriers (electrons) stored in die base region. When the transistor is driven<br />

into saturation, bodi junctions are forward biased. Thus a large concentration of electrons builds<br />

up in die base. Until diese electrons have been removed from the base, forward, current<br />

continues to flow across die junctions. Notice that the base current actually reverses directions at<br />

die end of die input pulse. This is due to stored charge flowing out of the base terminal.<br />

At about / = 0.52 p.s, most of die excess charge in the base has been removed and die collector<br />

current begins to fall as shown in Figure 5.48b, causing die output voltage to rise (Figure 5.48a).<br />

However, die output voltage rises gradually because of die junction capacitances. The transistor<br />

returns to die cutoff state at / > lu.s.<br />

The circuit of Figure 5.47 acts as a logic inverter. When die input is low, die output eventually<br />

becomes high. Similarly, when the input is high, the output eventually becomes low. Because of<br />

the charge storage effects, changes in the output do not occur immediately when die input<br />

changes. Of course, in most applications it is desirable for the switching delays to be as short as<br />

possible Several aspects of die device construction influence the switching speed. For example,<br />

reducing junction areas can reduce die device capadtances. Doping levels and junction grading<br />

also affect junction capadtances. A tiunner base region leads to quicker diffusion of minority<br />

charge carriers out of die base region. Selected impurities can be used to reduce die minority<br />

carrier lifetime.<br />

Often, BJT data sheets give specifications for switching time intervals for test circuits similar to<br />

die RTL inverter. We define die start of a logic transition as die point at which 10% of die<br />

voltage change has occurred. For example, the start of die leading edge of the 3-V input pulse is<br />

die point at which die input pulse reaches 0.3 V. Similady, die start of the leading edge of die<br />

output pulse is die point at which % has fallen to 2.7 V. These start points are labeled in Fig. 5.49.<br />

Similady, we define die end of a logic transition as die point for which 90% of the voltage<br />

(current) change has occurred.<br />

-179-


Analog Electronics /BJT Circuits<br />

4.6U-T ,<br />

2.8U.<br />

Start of leading edge of Vo<br />

•ts-<br />

•»-N<br />

if<br />

A<br />

2.7V<br />

/ End of<br />

/ trailing<br />

/ edge of Vo<br />

End of<br />

leading edge of Vo /<br />

z=*<br />

0.3V<br />

[Start of leading edge of Vin<br />

— i — - -<br />

h<br />

200ns<br />

400ns<br />

au(l) ou(3)<br />

Tine<br />

r -1 " "I<br />

660ns<br />

Figure~5.49 Waveforms illustrating turn-on and turn-off rimes<br />

Data sheets for BJTs intended for switching applications often specify the following switching<br />

intervals:<br />

• t d is the delay time measured from start of the input leading edge to the start of the output<br />

leading edge.<br />

• t r is the rise time, measured from the start point to the end point of the leading edge of the<br />

output pulse. Notice that the rise time is defined for the leading edge of the output pulse even<br />

though this is the negative-going edge.<br />

• /, is the storage time, measured from the start point on the trailing edge of the input pulse to<br />

the start point on the trailing edge of the output pulse.<br />

• ^is the fall time measured between the start and end points of the trailing edge of the output<br />

pulse.<br />

Turn-on and turn-off times, respectively t on =t d +t r and t 0 ff = f, +/y, are given on transistors<br />

data sheets. Typically, the switching times are much longer for power transistors than for smallsignal<br />

devices. This is due to larger junction capacitances and charge stored in a larger volume of<br />

the base region.<br />

C s =20pF<br />

Vaf*W<br />

Schottky-claapad RTL invartar<br />

Vin 1 0 FML(0 0 O.lu 0 O.llu 3 0.3u 3 0.31u 0)<br />

RB 1 2 5K<br />

Ca 1 2 20pF<br />

Ql 3 2 0 Q2N2222A<br />

Dl 2 3 DSH<br />

.MODIL DSH D(I«-l«-8 CJO-2pF VJ-0.6 M-0.5<br />

.LIB EVAL.LIB<br />

A (v 2N2222A RC 4 3 2k<br />

^ L Vcc 4 03V<br />

.TRAM O.lu lu 0 In<br />

. «nd<br />

Figure 5.50 RTL inverter with speed-up capacitor and Schottky clamp diode<br />

TT-0)<br />

-180-


Analog Electronics / BIT Circuits<br />

Techniques for speeding-up the transitions of the RTL inverter are of much interest in many<br />

applications. Two such techniques are illustrated in Figure 5.50. First, a speed-up capacitor has<br />

been added in parallel with die base resistor Rj. Second, a Schottky clamp diode has been<br />

added between the base and the collector terminal. Figure 5.51 shows waveforms generated by<br />

the PSpice program shown in Figure 5.50.<br />

*.4n<br />

fine<br />

«.tes<br />

Figure 5.51 Voltage waveforms for the Schottky-clamped RTL inverter<br />

Notice that die switching times for diis circuit are much smaller than for the simple RTL inverter<br />

of Figure 5.37. The speed-up capacitor couples the leading edge of the input pulse to the base.<br />

The voltage across the capacitor cannot change instantaneously. Thus the rapid increase in the<br />

input pulse causes a rapid increase in v BE . Thus the transistor is forced to turn on very quickly.<br />

Moreover, without the speed-up capacitor, the junction capacitances must be charged through<br />

Rg, and v BE rises more gradually.<br />

The Schottky diode prevents the transistor from entering the saturation region. When the<br />

collector voltage reaches approximately 0.4 V, the diode conducts and reduces the portion of the<br />

base current available to the transistor. Thus the output voltage is not allowed to fall below 0.4V<br />

and the transistor remains in the active region. This greatly reduces the concentration of electrons<br />

in the base region.<br />

The important point for the circuit designer is that if fast switching is important, the circuit should be<br />

designed so that the transistors do not enter saturation.<br />

5.16 Summary<br />

The bipolar junction transistor is a solid-state device consisting of two pn junctions fabricated in<br />

close proximity on single-crystal semiconductor. A large-signal, static, nonlinear Ebers-Moll<br />

model, embodied in an equivalent circuit and a pair of corresponding simultaneous equations,<br />

help us understand the transistor in terms of the theory of pn junctions. For small-signal linear<br />

applications, we use external dc sources to bias the transistor in its forward active state. For this<br />

state, the Ebers-Moll predicts the characteristic curves and linear equations of a currentcontrolled<br />

current source. This leads to a simple linear equivalent circuit that we use to predict<br />

how die transistor interacts with other circuit elements. To do the ac circuit analysis we replace<br />

me transistor with its equivalent circuit. Parameters of this circuit depend on the device jg-point<br />

that results from the particular bias. The BJT has two other states of major importance:<br />

saturation, in which the transistor resembles a closed switch, and cut-off, where the transistor<br />

-181-


Analog Electronics / BJT Circuits<br />

represents an open switch. The third state, inverse active, is of little practical importance. Each of<br />

die four states corresponds to operation in a particular region of the transistor output<br />

characteristics, and each has a circuit model, related in an obvious way to the characteristic<br />

curves. For amplifier applications the BJT is operated in the active state. Switching applications<br />

make use of the cut-off and saturation states. A BJT operating in the active state provides a<br />

collector current ic = isexp(J v BE \ /F T ). The base current is i B = ; c /p, and the emitter current i E =<br />

i c + i B . Also, i c = ai E and thus P=a/(l-a) and 50 V. For i E = 0, the breakdown voltage is less than<br />

BVCM- In the common-emitter configuration the breakdown voltage specified is BV^g, which is<br />

about half BV CBO . The emitter-base junction breaks down at the reverse bias of 6-8 V. There are<br />

also dynamic transistor limitations embodied in the depletion and diffusion capacitances of the<br />

junctions and between the collector and the substrate. Except for minor differences they are the<br />

same nonlinear capacitances we encounter in diodes. SPICE transistor models enable us to<br />

simulate both static and dynamic behavior of transistors, including all of the major nonlinearities.<br />

We generally use our simplified and more intuitive transistor concept to design, and then follow<br />

up with accurate computer simulations that include the second-order effects we consciously<br />

ignored to make our initial design work tractable. We examine and evaluate the simulation results<br />

in terms of our simple conceptual ideas and then redesign, if necessary. The dc analysis of<br />

transistor amplifiers is gready simplified by assuming diat | V BE \ = 0.7 V. To operate as a linear<br />

amplifier, the BJT is biased in the active region and the signal v h is kept small (v k « K T ). Bias<br />

design seeks to establish a dc collector current that is as independent of the value of p as<br />

possible. For small signals, the BJT functions as a linear voltage-controlled current source with a<br />

transconductance &n - IJ Vv Th e inp ut resistance between base and emitter, looking into the<br />

base is r„ = P/#n. For a common-emitter configuration, a high voltage gain and reasonable input<br />

impedance are obtained, but the high-frequency response is limited. Including an unbypassed<br />

resistance in the emitter lead can increase the input resistance of a common-emitter amplifier. In<br />

the common-base configuration, a high voltage gain (from emitter to collector) and an excellent<br />

high-frequency response can be obtained, but the input resistance is low. The CB amplifier is<br />

useful as a current buffer. In the emitter follower (common-collector amplifier) the voltage gain<br />

is less than unity, the input resistance is very high, and the output resistance is very low. The<br />

circuit is useful as a voltage buffer.<br />

-182-


Analog Electronics / Feedback Circuits<br />

6 Feedback Circuits<br />

Feedback consists of returning part of the output of a circuit or system to its input In an<br />

amplifier with negative feedback, a portion of the output signal is returned in opposition to the<br />

original input signal. In positive feedback, the feedback signal aids the original input Usually, in<br />

amplifiers, negative feedback is more useful than the positive feedback. However, the positive<br />

feedback is useful in the design of oscillators, which are considered later. The advantages of<br />

adding a negative feedback to an amplifier are<br />

• reduced sensitivity to parameter variations<br />

• increased bandwidth<br />

• reduced nonlinear distortion<br />

• improved input and output resistances<br />

One disadvantage is the reduction of gain; however, gain is easy to obtain with modern circuits,<br />

so this is a small price to pay. The other disadvantage, a possibility of unwanted oscillations or<br />

instability, is more serious, for oscillations make the circuit worthless as an amplifier.<br />

Fig. 6.1 shows the general structure of & feedback amplifier. A nonfeedback amplifier with gain A<br />

(the A circuit) delivers an output signal x, = Ax, to an external load. Instead of using signal x, as<br />

the amplifier input, a feedback circuit (p circuit) produces a feedback signal x f = fix, to form the<br />

actual input x> The feedback factor, P, used extensively in this chapter is unrelated to the bipolar transistor beta.<br />

Figure 6.1 Feedback amplifier structure<br />

Feedback theory applies to all four amplifier types named in Chapter 2.1,-but each type implies a<br />

different interpretation of Fig. 6.1. For a voltage amplifier, A denotes voltage gain and x e and x,<br />

are voltages. The external source must be a voltage source x s (ideally having zero internal<br />

resistance) and there is a voltage subtraction at the input. The feedback factor P = x/x. is a<br />

voltage gain (or attenuation) in this case.<br />

For a current amplifier, A denotes current gain. The input comes from a source of current x s<br />

(ideally having infinite internal resistance) and the current subtraction is required at the input.<br />

The p denotes the current gain of the feedback circuits.<br />

In a transconductance amplifier, A is transconductance gain with dimensions of amperes/volt.<br />

Thus output x, is a current and the source x t is the voltage source; the subtraction at the input is<br />

the voltage subtraction. For this case, P has a dimension of volts/ampere, or transresistance.<br />

Finally, the transresistance amplifier has transresistance gain with dimensions of volts/ampere.<br />

Thus output x c is a voltage and the external input source is the current source. The subtraction is<br />

a current subtraction. For this case, p has a dimension of transconductance.<br />

From Fig. 6.1 one can write<br />

-183-


Analog Electronics / Feedback Circuits<br />

x 0 = Axi = A(x s -x/) = A(x s - p* 0 )<br />

Solving for the gain of the feedback amplifier, Aj- = x 0 I x s , gives<br />

which applies to all four amplifier classes.<br />

Reduction of Gain. For midrange frequencies, both A and p are real quantities having the same<br />

algebraic sign. Equation (6.1) shows one of the prices we pay for negative feedback, a gain<br />

reduction. For example, suppose A~\W and the fraction of x, that is fed back to the input is<br />

P=0.01. Then the resulting feedback amplifier has gain of only<br />

10000 nn<br />

Usually, as in this illustration, ^4P»1, and (6.1) gives an important approximation<br />

A f =- (6-2)<br />

P<br />

For a voltage amplifier, common practice is to obtain P from a resistor network of lowtemperature<br />

coefficient, such as from a voltage divider, giving<br />

_ 1 R x +R 2 . %<br />

In an IC realization, such a P is easily controlled to within 1% or so by masking tolerances, and<br />

the temperature variations cancel over a wide temperature range. Thus Afis quite constant, even<br />

though A might vary greatly with temperature or other factors.<br />

Improvement Factor. The denominator of (6.1), 1+^P, is of considerable importance in<br />

feedback theory. This factor by which the gain is reduced also turns out to be the degree of<br />

improvement effected by introducing feedback. Ideally, parameter sensitivity, distortion,<br />

bandwidth, input resistance and output resistance all improve by \+A$ when one adds feedback.<br />

Thus the quantity is referred to as the improvement factor.<br />

6.1 Effects on Sensitivity, Bandwidth and Distortion<br />

We will first develop an intuitive notion of how the feedback works. Suppose in Fig. 6.1, while<br />

the input x t is held constant, output x, decreases in amplitude for any reason. Being proportional<br />

to *•„ feedback signal xyalso decreases, thereby increasing x ; to compensate, at least partially, for<br />

the original change. A complementary scenario opposes increases in x,. Thus negative feedback<br />

opposes any change in x, caused by an event unrelated to x> If the cause is variation in an<br />

internal parameter/), we say that feedback has reduced the sensitivity. If it is loss of gain at either<br />

low or high frequencies because of amplifier capacitances, we say feedback has improved the<br />

frequency response. If x, changes in a manner not proportional to x s because of nonlinearities<br />

within the amplifier, we say feedback has reduced distortion. We now examine the details to<br />

obtain a more quantitative understanding to use in analysis and design.<br />

6.1.1 Effect of Feedback on Sensitivity<br />

If the sensitivity of gain A to parameter/) is Sff, adding negative feedback results in an amplifier<br />

of gain Aj with lower sensitivity; specifically,<br />

-184-


Analog Electronics /Feedback Circuits<br />

To prove this, we need only use the definition<br />

s*f _ P M f _ P d[AI{\ + AP)}<br />

p<br />

Af dp Af dp<br />

Since the circuits we design employ feedback p that is independent of p, the derivative gives<br />

A f<br />

(1 + AP){dA I dp) - AfiidA I dp)<br />

(1 + A0) 2 P<br />

A f<br />

A/3 dA<br />

(1 + AP) (i + AJJ) 2 J dp<br />

Substituting for A, from (6.1) gives an equation that is equivalent to (6.3).<br />

Exercise 6.1 A nonfeedback amplifier with A = 347 contains a critical resistor R. By laboratory<br />

measurements we determine that SR =1.12. After adding negative feedback to reduce the<br />

sensitivity, we find the amplifier gain is 24. Find the gain sensitivity of the feedback amplifier to<br />

changes in K Ans. 0.077.<br />

Example 6.1 We have an amplifier with a gain of 800. Sensitivity of gain to temperature change<br />

is 0.1; however, our specification requires temperature sensitivity to be no more than 0.001.<br />

Investigate the possibility of using negative feedback to bring the temperature sensitivity within<br />

specification.<br />

Solution. From (6.3), we need an improvement factor of<br />

Since A = 800, P must satisfy<br />

1 + 800^=100<br />

or P = 0.124. The resulting feedback amplifier meets the temperature-sensitivity specification;<br />

however, its gain is only 800/100 = 8. +<br />

Sensitivity of Cascaded Feedback Amplifiers. Example 6.1 shows that we can improve<br />

sensitivity by sacrificing gain. Because specifications often involve both gain and sensitivity, it is<br />

interestirg to see if we can come out ahead by using a cascade'of low-gain, low-sensitivity<br />

feedback amplifiers.<br />

Consider a cascade of n feedback amplifiers, each of gain Af. We denote the overall gain by G =<br />

(A)", and now need to find the sensitivity of G to parameter/). By definition,<br />

SG EdG p ,dAf_<br />

p<br />

G dp G v J' dp<br />

Substituting G = (A)* gives<br />

cG P „,, sn-ldA* m P dA f<br />

Sp^niA/) -d7 =n^Hp-<br />

We conclude that<br />

Sf=nSp<br />

which shows that sensitivity increases only in proportion to n as we cascade stages, whereas G<br />

varies as die »-th power of gain. This suggests that we should be able to be better off using<br />

feedback. Table 6.1 illustrates the point using numbers from Example 6.1. Cascading feedback<br />

amplifiers is obviously an effective strategy.<br />

-185-


Analog Electronics / Feedback Circuits<br />

TABLE 6.1 GAIN AND SENSITIVITY OF CASCADED FEEDBACK AMPLIFIERS<br />

6.1.2 Effect of Feedback on Bandwidth<br />

n G<br />

^P<br />

1 ' 8 0.001<br />

2 64 0.002<br />

3 512 0.003<br />

4 4096 0.004<br />

Upper Half-Power Frequency. Consider using feedback to increase the bandwidth of an<br />

amplifier described by<br />

A = A(co) = A 0 , *" (6.4)<br />

where A e is the midband gain. Substituting this expression into (6.1) gives<br />

a> H /(ja) + a> H )<br />

Af<br />

or<br />

— A t<br />

f °l + A 0 [a> H /(j6> + o> H )y3<br />

= 4<br />

A f = A of<br />

where<br />

° ja> + H p<br />

+ a, H (\ + A o P)\ + A o 0<br />

a>Hf<br />

(6.5)<br />

J6) + a> Hf<br />

A n t =<br />

of = l + A B<br />

and<br />

a W = aH ® + A °®'<br />

Comparing (6.5) and (6.4) one can see that the feedback amplifier's gain function has the same<br />

general form as the original; however, its upper half-power frequency is higher by (l+A$) and its<br />

midrange gain is lower by the same factor. Figure 6.2a compares the two frequency response<br />

curves. Notice the original and final gain-bandwidth products, Afi) H and A^Q) H j. We see that the<br />

gain-bandwidth product is preserved when one adds feedback. We conclude that for a one-pole<br />

amplifier described by (6.4), negative feedback facilitates a direct trade-off of gain for bandwidth<br />

as in Fig. 6.2a. Amplifiers with more complex gain functions do not give an exact trade-off;<br />

however one can expect a dominant high-frequency pole to approximate such a trade-off.<br />

Gain (dB)<br />

Gain (dB)<br />


Analog Electronics / Feedback Circuits<br />

Lower Half-Power Frequency. Consider a nonfeedback amplifier described by<br />

id)<br />

A = A(a>) = A 0 / (6.6)<br />

Substituting into (6.1) gives<br />

A =A ja>l(J(Q + a>L)<br />

f °\ + A o [ja>/U L )]fi<br />

Algebraic manipulations similar to those for the high frequency case lead to<br />

id)<br />

A f = A of -H (6.7)<br />

J<br />

° J ja) + 80, f L < 50 Hz,<br />

f H > 15 kHz, and sensitivity to power supply changes < 0.2. We have an amplifier with gain =<br />

1000,^ = 400 Hz,f H — 9 kHz and its sensitivity to changes in supply voltage is 1.5. Determine<br />

whether we can meet the specifications by adding feedback to the existing amplifier. If so, find an<br />

acceptal e value of P and give final specifications for the feedback amplifier.<br />

Solution. Assume that a pair of dominant poles characterizes the frequency response.<br />

Considering each specification individually gives the following requirements:<br />

1000<br />

gain: -—,^,,^80<br />

6<br />

1 +1000^<br />

400<br />

lower -3 dB frequency:<br />

M }<br />

-—, nnn •- < 50/fe<br />

1 + 1000/<br />

upper -3 dB frequency: 9000(1 + 1000/ff) £ 15000/fe<br />

sensitivity: -—.„„„„ ^ 0.2<br />

y<br />

1 + 1000^<br />

These lead, respectively, to the requirements l+10 3 p < 12.5, 1+HPP > 8, l+10 3 p > 1.67, and<br />

1+10 3 P > 7.5. Sorting out these inequalities shows that any P that satisfies 8 < 1+10^ < 12.5 is<br />

acceptable. To leave room for error, choose 1+10 3 P = 10, which requires P = 0.009. Then the<br />

final amplifier will have gain = 100,^ = 40 Hz,^ = 90 kHz and supply sensitivity of 0.15, all<br />

within specifications.<br />

-187-


Analog Electronics / Feedback Circuits<br />

6.1.3 Effect of Feedback on Nonlinear Distortion<br />

Nonlinear distortion is normally not a problem in the small-signal amplifiers. However, in power<br />

amplifiers, output signal must be large enough to develop a specified amount of power in the<br />

load. This often requires signal swings large enough to traverse a significant portion of the<br />

amplifier's nonlinear transfer characteristic. When this occurs, the output signal is the amplified<br />

input signal, plus additive harmonic distortion components that arise within the amplifier itself.<br />

Amazingly, negative feedback can be used to feed these distortion components back into the<br />

input in such a way that they subtract from themselves. The feedback also causes some signal to<br />

subtract from itself; however, this is the gain reduction we expect with negative feedback, and we<br />

can compensate for it by adding a preamplifier. The bottom line is that negative feedback reduces<br />

internally generated distortions by (\+A$). The following quasi-numerical development shows<br />

how it works.<br />

XJ(0=0.001COS(


Analog Electronics / Feedback Circuits<br />

x o (t) = d(t) + A[x s (t)-0 Xo (t)]<br />

or<br />

d(t)<br />

\ + A0 \ + A0 MO sy<br />

Using P = 0.01, as an example, gives<br />

x 0 (0 = 0.0099 COS(GJT) + 99x s (r) (6.8)<br />

showing that distortion is reduced to 1 % of its former value. Notice, however, that the new<br />

input xj(t} is multiplied by only 99 in (6.8) instead of the original 10 4 . Because specifications<br />

require 10cos(©i) at the output, from Eq. (6.8) the input to the feedback amplifier must satisfy<br />

99x s (0 = 10cos(fi#)<br />

or x,= 0.101cos(©/), showing that the feedback amplifier requires a larger input voltage that the<br />

nonfeedback amplifier for the same output (The corresponding transfer characteristic of the<br />

feedback amplifier, x, versus x„ is much more linear than the original curve, but has lower slope.)<br />

Since only 0.001cos(©/) is available for us for our io^ut signal, we must add a preamplifier of gain<br />

Ap at the input that satisfies<br />

0.001^=0.101<br />

or<br />

4, = 101.<br />

Figure 6.3c summarizes the design, both in general notation and in the numerical values used as<br />

examples. For this strategy to succeed, we must construct a preamplifier that does not produce<br />

distortion. The key observation is that the signal amplitude required at the preamplifier output is<br />

much smaller than the specified amplifier output voltage. That is, it is much easier to produce a<br />

distortion-free output voltage of 0.101 V than 10 V.<br />

Vcc = +15V<br />

V<br />

K £ £ = -15V<br />

Figure 6.4 Nonlinear class-B power amplifier<br />

Example 6.3: Crossover distortion. Distortion is mainly a problem in power amplifiers. A<br />

simplified example of an audio-amplifier output stage is shown in Fig. 6.4. Notice that if v s = 0,<br />

both transistors are in cutoff since there is no forward bias of the base-emitter junctions. In fact,<br />

neither transistor conducts until v, swings outside the range from -0.6 V to 0.6 V (for typical<br />

silicon power transistors). With both transistors cut off, the output voltage is zero. As v, swings<br />

higher than 0.6 V, the npn transistor J2, turns on and supplies current to the load. In this case the<br />

output voltage is given by, approximately,<br />

v 0 = v 5 - 0.6 for v s > 0.6<br />

If v s is less than -0.6 V, the npn is off and the pnp transistor Q^ is in the active region. Then we<br />

have<br />

v 0 =v s + 0.6 for v,


Analog Electronics / Feedback Circuits<br />

The transfer characteristic for the amplifier is shown in Fig. 6.5. Notice the nonlinearity in the<br />

region around v t = 0. This nonlinearity causes ctossover distortion when conduction is changing<br />

from one transistor to the other. Also, notice that the voltage gain, which is the slope of the<br />

transfer characteristic, is approximately unity (except in the region around zero input).<br />

5.W<br />

-S.W +<br />

-6.W<br />

-4.MI<br />

Figure 6.5 Transfer characteristic for the amplifier of Fig. 6.4<br />

This circuit is an example of a class B amplifier, in which each device conducts for<br />

approximately half of the signal cycle. Figure 6.6 shows the class B output stage driven by a<br />

differential amplifier that has a differential gain of 1000. The feedback network consisting of R,<br />

and R 2 returns part of the output voltage to the inverting input of the differential amplifier.<br />

Normally the switch would be in position JB, so the output voltage across the load is fed back.<br />

However, we will also analyze the circuit with the switch in position A to illustrate the crossover<br />

distortion of the output stage.<br />

The feedback ratio P is given by<br />

V *2<br />

P =<br />

Ri +R-, = 0.1<br />

"o -4 ••• A 2<br />

Since the gain of the differential amplifier is 1000 and the gain of the class B stage is<br />

approximately unity, the overall open-loop gain isA= 1000. Thus we have A$ = 100, which is<br />

much larger than unity. Consequendy, we expect to 6ndA/= 1/P = 10.<br />

RL=&n<br />

00<br />

F££=-15V<br />

-190-


Analog Electronics /Feedback Circuits<br />

Figure 6.6 Class B power amplifier with feedback:<br />

(a) circuit diagram, (b) model for die differential amplifier<br />

The differential amplifier provides the means for subtracting the feedback signal v } from the<br />

source voltage. We use the circuit model for the differential amplifier shown in Fig. 6.6b. The<br />

SPICE program to analyze the circuit is listed below.<br />

Class B Feedback Example<br />

Vs 1 0 sin(0 0.2 1000)<br />

82 4 0 Ik<br />

Rl 2 4 9k; change first node to 3 for switch at B<br />

EA 2 0 1 4 1000; voltage-controlled voltage source<br />

Rin 1 4 1MEG<br />

Ql 5 2 3 npnpower<br />

Q2 6 2 3 pnppower<br />

RL 3 0 8.0<br />

Vcc 5 0 15V<br />

Vee 6 0 -15V<br />

.model npnpower npn(bf=150 Is=le-12)<br />

.model pnppower pnp(bf=150 Is=le-12)<br />

.tran lOOus 2ms 0 5us<br />

.end<br />

After executing this program, we request plots of the drive voltage V(2) at the bases of J2, and Q 2<br />

as well as the output voltage v t — V(3). The result is shown in Fig. 6.7a. In this case, the switch is<br />

in position A, so the nonlinearity of the output stage is not included in the feedback loop. The<br />

output signal waveform thus deviates from an ideal, distortion-free shape and, consequendy, a<br />

significant distortion signal d(t) is a part of the output, as is also shown in Fig. 6.7a. The base<br />

drive is sinusoidal, but the output demonstrates considerable crossover distortion.<br />

-191-


Analog Electronics / Feedback Circuits<br />

Tlaa<br />

(b) Waveforms for the switch in position B<br />

Figure 6.7 Waveforms for the circuit of Fig. 6.6<br />

Then we change the program to place the switch in position B, obtaining the results shown in<br />

Fig. 6.7b. In this case the output voltage is almost free of distortion, d{t) = 0. Notice that the base<br />

drive voltage V(2) has been predistorted to compensate for the nonlinearity of the output stage.<br />

Also notice that, as expected, the output voltage is Aj = 10 times larger that the source voltage<br />

signal.<br />

Exercise 6.3 Suppose that we need to change the amplifier of Fig. 6.6, so that the gain Aj is<br />

approximately 20. What changes do you suggest Include component values.<br />

Exercise 6.4 Change the resistor values in Fig. 6.6 to R 2 = lOfi and R, = 9990Q. What is the<br />

approximate value of A$ Use a Spice program to find the output waveform for the switch in<br />

position B and v s — 0.004sin(2000rc/). Is the feedback effective in reducing distortion in this case<br />

Explain.<br />

6.1.4 Effect of Feedback on Noise<br />

An undesirable aspect of amplifiers is that they add unwanted noise to the desired signal. Sources<br />

of this noise include power-supply hum, coupling of signals from other circuits, and thermally<br />

generated noise in resistors. Another noise is shot noise caused because current flow is not<br />

continuous; instead, charge is carried in discrete quantities by individual electrons. Still another<br />

source is microphone noise, which is an electrical signal arising from vibration of circuit<br />

mechanical components.<br />

Some source of noise can, in principle, be eliminated. For example, power-supply hum can be<br />

reduced by additional filter circuitry in the power supply. However, some of the noise sources,<br />

such as thermal and shot noise, stem from the basic natural processes and cannot be totally<br />

eliminated. Thus all amplifiers add noise, but some amplifiers are much worse than others are. In<br />

this section we wish to show that feedback can, under certain circumstances, reduce noise.<br />

If d(t) introduced in Section 6.2.3 represents noise arising within the nonfeedback amplifier,<br />

negative feedback reduces output noise just as it reduces distortion; however, it can be shown<br />

that adding negative feedback results in no improvement in the input signal-to-noise-ratio of an<br />

amplifier. Since this ratio is usually the critical factor, adding negative feedback is not a generally<br />

effective noise reduction strategy.<br />

-192-


Analog Electronics / Feedback Circuits<br />

Figure 6.8 Model that accounts for the addition of noise in amplifiers<br />

The additive noise can be modeled as shown in Fig. 6.8. The amplifier gain is denoted as A,. To<br />

quantify noise performance, engineers use the signal-to-noise tatio, which is the desired signal<br />

power delivered to the load divided by the noise power. We denote the rms values of the signal<br />

and noise by X, and X,. The rms signal delivered to the load in Fig. 6.8 is A,X S and the rms noise<br />

is A,X M . If the signals are voltages, the powers delivered to the load are<br />

Ps =<br />

and<br />

(<br />

Ps = (Ax ny<br />

^ (6.9)<br />

R,<br />

The signal-to-noise ratio is given by<br />

(6.10)<br />

Equation (6.11) also applies if X s and X, are currents.]<br />

(6.11)<br />

Figure 6.8 Model that accounts for the addition of noise-in amplifiers<br />

Now consider the feedback amplifier shown in Figure 6.9. X, is the rms noise of amplifier A,.<br />

Amplifier A 2 has been added to increase the open-loop gain of the amplifier. Amplifier A 2 is<br />

assumed to be noise-free. This is a reasonable assumption if we have a situation in which<br />

amplifier A, is very noisy and amplifier A 2 is well designed, so its noise is very small. For<br />

example, A, can be a very high power amplifier with a great deal of power-supply hum, but A 2 is<br />

supplied with a well-filtered power. (Perhaps the designer is trying to be economical by using less<br />

filtering for the power to amplifier^!,.)<br />

We analyze the system shown in Figure 6.9 to find an expression for the signal-to-noise ratio. We<br />

can write<br />

x 2 {t) = x s {t)-px 0 {t) (6.12)<br />

x l (t) = A 2 x 2 (t) + x n (t) (6.13)<br />

x 0 (t) = A lXl (t) (6.14)<br />

Substitution of Equation (6.12) into (6.13) and the result into (6.14) results in<br />

x 0 {t) = A x {A 2 [x s (t) - Px 0 {t)\ + x n {t)} (6.15)<br />

Solving for x,[t), we find that<br />

-193-


Analog Electronics / Feedback Circuits<br />


Analog Electronics /Feedback Circuits<br />

The units of p are the inverse of the units of the gain for each type of feedback. Refer to Figure<br />

6.9b, as an example. Notice that the units of the transconductance gain G m are Siemens. Also, we<br />

see that Vf=$i c . Therefore, p is a transresistance parameter with units of ohms.<br />

Similarly, for parallel voltage feedback, the gain parameter is a transresistance, and the feedback<br />

ratio p is a transconductance. For series voltage feedback A = A r which is unidess, and P is also<br />

unidess. Finally, for parallel current feedback, A = A t and P are both unitless.<br />

The effect of each type of feedback on input and output impedance of an amplifier is different.<br />

In design, we select the type of feedback in accordance with the design objectives.<br />

(b) Series current feedback<br />

(c) Parallel voltage feedback<br />

(d) Parallel current feedback<br />

Figure 6.9 Types of feedback<br />

-195-


Analog Electronics / Feedback Circuits<br />

6.3 Effect of Feedback Types on Input and Output Impedance<br />

Now we examine the effect of series feedback on input impedance. The model for our discussion<br />

is shown in Figure 6.10. The output signal x, is sampled by the feedback network, which<br />

produces a feedback voltage signal v f = fix, connected in series with the source and the input<br />

terminals of the amplifier. The original (before the feedback is added) input impedance of the<br />

amplifier is R,. The input impedance of the amplifier with feedback is<br />

%=7" (6-18)<br />

Writing a voltage equation around the loop in Figure 6.10, we obtain<br />

v, = *,-/, + v f (6.19)<br />

But «y=P^ so we have<br />

v s = R,i s +0K o (6.20)<br />

Also, the input voltage is given by<br />

Vi = Rii s (6.21)<br />

and the output is given by<br />

x 0 = Av t (6.22)<br />

where A=A r is a voltage gain if x,=v„ or A—G m is a transconductance gain if die output is a<br />

current x=i,. Substituting Equation (6.21) into (6.22) and the result into (6.20), we obtain<br />

v^Riis + A/SR^ (6.23)<br />

which can be solved for the input impedance with feedback<br />

R if =f = R i (l + Afi) (6.24)<br />

Recall that for negative feedback the factor (\+A$)<br />

feedback increases input impedance.<br />

is larger than unity. Thus negative series<br />

Load<br />

Figure 6.10 Model for analysis of the effect of series feedback on input impedance<br />

Load<br />

Figure 6.11 Model for analysis of the effect of parallel feedback on input impedance<br />

Next, we consider the effect of parallel feedback on the input impedance. The model is shown in<br />

Figure 6.11. It can be shown that the input impedance in this case is<br />

-O-<br />

-196-


Analog Electronics / Feedback Circuits<br />

Ri<br />

Thus negative parallel feedback reduces input impedance.<br />

Exercise 6.5 Derive Equation (6.25).<br />

To find the output impedance of an amplifier, we turn off die input source, remove the load, and<br />

look back into the output terminals. A model for the voltage feedback amplifier with these<br />

changes is shown in Figure 6.12. A test voltage source v M has replaced the load at the output<br />

terminals of the feedback amplifier. The output impedance with feedback is<br />

v<br />

R test<br />

of =<br />

(6.26)<br />

hest<br />

To simplify our analysis, we assume that the input impedance of the feedback network is infinite.<br />

Thus the feedback network does not load the amplifier output.<br />

Xs=0 +<br />

'Jest<br />

v test<br />

Figure 6.12 Model for the analysis of output impedance widi voltage feedback<br />

The output circuit of the amplifier is modeled by a controlled voltage source with gain parameter<br />

A^ The subscripts of the gain parameter indicate that it is the open-circuit amplifier gain. If x,=f,<br />

we have series voltage feedback and A K = A.,^. On the other hand, if x~ij, we have parallel<br />

voltage feedback and A K = R^ In any case, the resistance R, shown in Figure 6.12 is the output<br />

resistance of the amplifier before feedback.<br />

For die output loop of Figure 6.12 we can write<br />

v test ~ ^ohest + ^oc x i (6.27)<br />

However, we have<br />

*/ = -Potest (6-28)<br />

Substituting Equation (6.28) into (6.27) and solving for the output resistance with feedback, we<br />

have<br />

v<br />

R test<br />

R„<br />

of =<br />

(6.29)<br />

hest<br />

1 + fiA,<br />

oc<br />

Thus negative voltage feedback reduces the output impedance of an amplifier.<br />

Next, we consider the effect of current feedback on output resistance. The model for this analysis<br />

is shown in Figure 6.13. As before, the source signal x/is set to 2ero, the load is removed, and a<br />

test source is connected to the output terminals. The feedback network is assumed to have zero<br />

input impedance, so it produces no loading effects at the amplifier output.<br />

197


Analog Electronics / Feedback Circuits<br />

x=0 +<br />

^-40-^<br />

&==&, a ^"Irjf<br />

Figure 6.13 Model for the analysis of output impedance with current feedback<br />

The output of the amplifier is modeled by a controlled current source in parallel with the output<br />

resistance. The gain parameter A x has subscripts indicating that it is the gain of the amplifier with<br />

a short-circuited load. For parallel current feedback, the gain is the short-circuit current gain A x<br />

= A isf For series current feedback, the gain is the short-circuit transconductance gain A x = G au ,<br />

For the system of Figure 6.13 we can show that<br />

R of = v test = R 0 (1 + J3A SC ) (6.30)<br />

hest<br />

Thus negative current feedback increases the output impedance of an amplifier.<br />

Exercise 6.6 Derive Equation (6.30).<br />

6.4 Summary of the Effects of Various Feedback Types<br />

We have seen that four types of feedback are possible. One effect of feedback is to stabilize and<br />

linearize gain. (i.e. Aj tends to be independent of A). However, the particular type of gain<br />

stabilized depends on the type of feedback. Table 6.1 shows the type of gain stabilized and<br />

linearized for each type of feedback.<br />

We have seen that (negative) series feedback increases input impedance, whereas parallel<br />

feedback reduces input impedance. If yip is very large, the input impedance tends toward either<br />

an open circuit or a short circuit. The formulas for input impedance are shown in Table 6.1.<br />

TABLE 6.1 EFFECTS OF FEEDBACK<br />

FEEDBACK<br />

TYPE x t x f<br />

'<br />

GAIN<br />

STABILIZED<br />

INPUT<br />

IMPEDANCE<br />

Series voltage t>, *>. A V-\ + A v 0 RiQ + A v /><br />

Series current », i.<br />

A<br />

OUTPUT<br />

IMPEDANCE<br />

Ro<br />

1 + AvocP<br />

G m<br />

mf ~\ + G m p Ri(l + G m )T> R 0 Q + G msc P)<br />

IDEAL<br />

AMPLIFIER<br />

Voltage<br />

Transconductance<br />

Parallel<br />

Ri Ro Transresistance<br />

Rm<br />

j<br />

voltage ', ».<br />

i + R m P 1 + RmocP<br />

Parallel<br />

Ai<br />

A<br />

Ri<br />

current *, '. »-l + A iP 1 + 4/ R o (\ + A isc 0) Current<br />

-198-


Analog Electronics /Feedback Circuits<br />

To reduce output impedance we could employ voltage feedback. On the other hand, to increase<br />

output impedance, we could choose current feedback. Of course, in making these statements, we<br />

assume negative feedback - the effect of positive feedback is the opposite. Table 6.1 also<br />

contains formulas for the output impedance for each of the four feedback types.<br />

We can summarize the effect of each type of feedback by stating that it tends to produce an ideal<br />

amplifier of a certain type. For example, series voltage feedback increases input impedance,<br />

reduces output impedance and stabilizes voltage gain. Thus series voltage feedback tends to<br />

produce an ideal voltage amplifier. As summarized in Table 6.1, similar statements can be made<br />

for the other feedback types.<br />

6.5 Practical Feedback Networks<br />

So far, we have modeled feedback networks as controlled sources. This approach simplified the<br />

analysis and allowed us to focus on the main effects of the various types of feedback. However,<br />

in practice we use simple networks of resistors (or in some cases resistors and capacitors). This<br />

components are available with precise and stable values (over time and with temperature<br />

changes) compared to the parameter values of active components (transistors). We employ<br />

negative feedback, so the amplifier characteristics depend mainly on the feedback network,<br />

thereby achieving amplifiers having precision and stability. Figure 6.14 shows examples of<br />

feedback amplifiers using practical resistive feedback networks.<br />

V f<br />

n<br />

(a) Series voltage, p = = — — (assuming U £ 0)<br />

v 0 R\ + R 2<br />

R 2<br />

f<br />

(b) Series voltage, / = -;— = Rf (assuming U = 0)<br />

-199-


Analog Electronics / Feedback Circuits<br />

l f<br />

n<br />

(c) Parallel voltage, p = — = — —— (assuming v,- 2 0)<br />

v o<br />

>L _<br />

l<br />

R f<br />

n if R\<br />

(d) Parallel current, p = T~ = ~ T 7T (assuming p,- = 0)<br />

l 0 R\ + K 2<br />

Figure 6.14 Examples of resistive feedback networks<br />

Notice that we have modeled the source as a voltage source for series feedback and as a current<br />

source for parallel feedback. This is consistent with Figure 6.9. The Thevenin model for the<br />

source is more natural for series feedback because the feedback voltage j^is subtracted from the<br />

source voltage v, in a series connection. The Norton model is more natural for parallel feedback<br />

because the feedback current ij\& subtracted from the source current i s in a parallel connection.<br />

Each of the feedback amplifiers shown in Figure 6.14 has negative feedback. For example, in the<br />

series voltage case shown in part (a) of the Figure, suppose that v s has a positive value. This<br />

results in a positive voltage at the noninverting input. The amplifier, in turn, produces a positive<br />

output voltage. The feedback network, composed of R, and R*, returns a fraction of the output<br />

voltage to the inverting input. This reduces the input voltage v f Thus the feedback signal acts in<br />

opposition to the original source signal, and we have negative feedback. If the inverting and<br />

noninverting input terminals were interchanged, positive feedback would result. A similar<br />

discussion applies to the remaining feedback amplifiers shown in Figure 6.14 b, c, and d.<br />

We can identify series feedback and parallel feedback by examination of the circuit configuration<br />

at the amplifier input. Study Figure 6.14a and b to verify that the signal source, the amplifierinput<br />

terminals and the output of the feedback network are in series. Also, verify the parallel<br />

connection for Figure 6.14c and d.<br />

To test for cutrent feedback, open-circuit the load so that the output current becomes zero. If<br />

the signal returned to the amplifier input by the feedback network becomes zero, the amplifier<br />

has current feedback.<br />

-200-


Analog Electronics / Feedback Circuits<br />

To test for voltage feedback, short-circuit the load so that the output voltage becomes zero. If<br />

the signal returned to the amplifier input by the feedback network becomes zero, the amplifier<br />

has voltage feedback. Verify that the types of feedback are correcdy labeled in Figure 6.14 by use<br />

of these tests.<br />

Exercise 6.7 For each of the circuits shown in Figure 6.15, identify the type of feedback present<br />

(negative-positive, series-parallel, and voltage-current). Determine the value of the feedback<br />

ratio, assuming zero input current and zero input voltage of the non-feedback amplifier, where<br />

appropriate. What type of ideal amplifier results if A$ is very large What is the gain of this ideal<br />

amplifier What value (0 or oo) do the input and output resistances approach Ans. (a) Negative<br />

series voltage, 3 = 1, ideal voltage amplifier, A 4 =l,K,-=°o,R, = 0<br />

(b) negative parallel current, P = 1, ideal current amplifier, A^= 1, R, = 0, R, = oo,<br />

(c) negative parallel voltage, P=-1/(3R), ideal transresistance amplifier, R^=-3R, R,=0, R„=0,<br />

(d) negative series current, P=R/2, ideal transconductance amplifier, G^ = 2/R, R,= oo, R = oo,<br />

(e) negative series voltage, p=l/12, ideal voltage amplifier, >4^= 12, R, = oo, R, - 0.<br />

-201-


Analog Electronics / Feedback Circuits<br />

6.6 Stability of Feedback Amplifiers<br />

In the above discussion, the nonfeedback amplifier (A circuit) was considered memoryless, for<br />

simplicity. Therefore, a zero phase shift between its output and input was assumed. Because of<br />

nonzero phase shift within any real amplifier, it is possible that a component of the feedback<br />

signal at some particular frequency/ is actually added rather than subtracted from die signal<br />

source. If this component is sufficiently large, the result is sustained oscillation at frequency/ -<br />

the circuit has become an oscillator or signal generator instead of an amplifier, and we say the<br />

circuit is unstable.<br />

AiS)<br />

Figure 6.16 Loop gain in a feedback amplifier<br />

For concrete illustration of how oscillations can occur, consider a series voltage circuit of Figure<br />

6.16, where sinusoidal voltages are indicated as phasors. With the switch in position 1, we have<br />

the feedback amplifier; position 2 gives the nonfeedback amplifier. Usually, P is a constant and<br />

does not contribute to the phase shift At midband frequencies, A((0) is real. Therefore, with die<br />

switch in position 1, the feedback phasor, Vf=\SA(


Analog Electronics / Feedback Circuits<br />

When this occurs, we say we have positive feedback at frequency/. It is possible for feedback<br />

to be negative for midrange components and positive for frequencies outside the -3dB passband<br />

of A(f) where additional phase shift occurs.<br />

With the switch in position 2 and sinusoidal excitation at frequency/, suppose in Equation (6.36)<br />

that in addition to ^(/ 0 ) = 180°, we also have<br />

Mfoi * i ( 6 - 38 ><br />

This amplitude condition leads to two interesting possibilities. Equality in Equation (6.38)<br />

implies<br />

v f (/) = V sm{l7rf 0 t +180°) = -V sm{2nf 0 t +180°) (6.39)<br />

When this condition is satisfied, we can use Vj to replace V f That is, once the circuit is running,<br />

we can flip the switch to position 1, turn off the external signal and use V { as the input signal.<br />

Thereafter, there would be a sustained sinusoidal output at frequency/. Circuits that operate like<br />

this are called non-self-starting oscillators.<br />

When the inequality occurs in Equation (6.38), V f is identical in frequency and phase to V s but<br />

greater in amplitude. In this case, the sinusoid of frequency/ increases in strength until circuit<br />

nonlinearities limit its amplitude. The result is a sustained oscillation with a periodic output<br />

waveform of fundamental frequency/. A circuit deliberately designed to operate in this way is<br />

called a self-starting oscillator. If such oscillations occur in an amplifier, we say the amplifier is<br />

unstable.<br />

Taking into account the complex-valued gain of the nonfeedback amplifier, the closed-loop gain<br />

of a feedback amplifier, as given by Equation (6.1), can be described by<br />

A{f)<br />

Af(f) = —^±L-!— (6.40)<br />

f U J<br />

K }<br />

l + Mf) QA(f)<br />

For a given frequency/, if pVl(/5 = -1> the closed-loop gain becomes infinite. This means that<br />

even a very small input signal may cause an infinite amplifier response. Such an amplifier is<br />

unstable and may produce oscillations with no input signal. The amplitude and phase conditions<br />

associated with oscillations both are related to the complex product<br />

L(f)=6A(f) (6.41)<br />

called the loop gain. The loop gain L0 is the total gain of the feedback loop from the amplifier<br />

input back to the point of signal subtraction, including all loading effects at input and output.<br />

Figure 6.16 shows that JL0 is the gain VJ V s when the switch is in position 2, that is with the<br />

feedback loop open.<br />

We now state the Nyquist stability criterion, a necessary and sufficient condition for feedback<br />

amplifiers to be unstable. If there exists any frequency/ such that the loop gain<br />

L(f 0 ) = jfl4(/ 0 ) = MZl80°, and M > 1 (6.42)<br />

then the amplifier is unstable, and there will be oscillations at/. Otherwise, the amplifier is stable.<br />

The justification for such a simple stability criterion is that random noise is present in every<br />

circuit with its power distributed over all frequencies - an infinity of tiny signal sources. For this<br />

reason, if there is any frequency whatsoever at which Equation (6.42) is satisfied, oscillations are<br />

inevitable even with no external signal applied to the circuit With an external signal applied, the<br />

oscillations simply superimpose upon it<br />

-203-


Analog Electronics / Feedback Circuits<br />

Gain Margin and Phase Margin. In design of feedback amplifiers, it is often helpful to<br />

consider Bode plots of me magnitude and phase of the loop gain $A(fi. For real and positive (3,<br />

its effect on the magnitude plot is simply to shift it vertically by 201og(P) and there is no effect on<br />

the phase plot. Thus the Bode plots of the loop gain $A(fi are the same as the Bode plots of the<br />

amplifier open-loop gain A(J) except for the vertical shift in the magnitude plot<br />

An example of the Bode plots for a BJT amplifier is shown in Figure 6.17. In considering the<br />

stability of an amplifier, we examine the Bode plot for the loop gain $A(fi to find the frequency<br />

f GM fot which the phase shift is 180°. If the magnitude of the loop gain is less than unity at this<br />

frequency, die amplifier is stable. On die other hand, if the loop gain magnitude is greater than<br />

unity, the amplifier is unstable.<br />

For a stable amplifier, the gain at/^ is less than unity in magnitude (it is negative when expressed<br />

in decibels). The amount mat the gain magnitude is below 0 dB is called die gain margin. The<br />

gain margin is illustrated in Figure 6.17. It can be shown that a gain margin of zero implies a pole<br />

lies on they'll) axis on the .r-plane. As gain margin becomes larger, the pole moves back into die<br />

left half of the .r-plane. In general, larger gain margin results in less ringing and faster decay of the<br />

transient response.<br />

Another measure of stability that can be obtained from the Bode plots is the phase margin.<br />

Phase margin is determined at the frequency f m for which the loop gain $A(f PA j) is unity in<br />

magnitude (i.e., 201og | fyAffpy) \ — 0 dB). The phase margin is the difference between me actual<br />

phase and 180°. This is also illustrated in Figure 6.17.<br />

As we noted earlier, we usually want to design feedback amplifiers to avoid ringing transient<br />

response and gain peaks in the frequency response. A. generally accepted rule of thumb is to design for a<br />

minimum gain margin qflOdB and a minimum phase margin of 45 °<br />

For the amplifier illustrated by Figure 6.17, the frequency f GM is approximately equal to 30.1<br />

MHz, and the gain margin equals to 21.7 dB. The phase margin can be determined at frequency<br />

f PM £ 3.1 MHz as being equal to 79°. Thus the example amplifier is stable and shows reasonable<br />

gain and phase margins.<br />

-204-


Analog Electronics / Feedback Circuits<br />

Often an amplifier is unstable for a given gain l/p\ which is especially true for multistage<br />

nonfeedback amplifiers that have complicated frequency response. We can make such amplifier<br />

stable by altering its open-loop gain curve A(J). Deliberately changing the frequency response of<br />

the nonfeedback amplifier to make a feedback amplifier stable is called frequency<br />

compensation. Some techniques have been developed for frequency compensation, such as pole<br />

addition used for operational amplifier design. However, the topic of frequency compensation is<br />

outside the scope of this textbook. The modern operational amplifiers, which are most often<br />

used as nonfeedback amplifiers, are internally compensated by adding an on-chip capacitor to<br />

introduce the dominant pole for compensation.<br />

Exercise 6.8 Use PSpice to find gain margin and phase margin for the uA741 operational<br />

amplifier. Hint Run the .ac analysis and use evailib library.<br />

6.7 Sinusoidal Oscillators<br />

Oscillators are intentionally unstable circuits that serve as sources of electrical waveforms. There<br />

are two broad classes of oscillators: sinusoidal oscillators, which produce sinusoidal waveforms,<br />

and relaxation oscillators, which produce triangular, or rectangular, waveforms. Both classes of<br />

oscillators are widely used for time bases in test and measurement equipment, and for signal<br />

processing in analogue and digital communication systems. Here we concentrate on sinusoidal<br />

oscillators.<br />

6.7.1 General Theory of Sinusoidal Oscillators<br />

A sinusoidal oscillator has three functional parts, a phase shifter to establish the frequency of<br />

oscillation, a gain circuit to compensate for energy losses in the phase shifter, and a limiter to<br />

control the amplitude of the oscillations. The gain circuit might be an operational amplifier or a<br />

transistor amplifier. The phase shifter is typically an RC or JJC circuit. The limiter might be a<br />

diode, a thermistor, or a variable-gain amplifier. In some oscillators the basic functions are<br />

combined rather than relegated to individual subcircuits. For example, the internal capacitances<br />

of the transistor that provides gain might contribute to the phase shifter, and inherent transistor<br />

nonlinearities often provide the limiting. Common to all oscillator circuits is instability, which is<br />

best understood in terms of positive feedback theory.<br />

Figure 6.18 General structure of a sinusoidal oscillator<br />

The voltage-shunt feedback structure of Fig. 6.18 describes many sinusoidal oscillators. A<br />

voltage amplifier with gain A((o) = VJV-, provides the gain, and a feedback network described<br />

by P(co) = VjfV, is the phase shifter. An oscillator representation, such as Fig. 6.18, differs in<br />

several ways from a feedback amplifier diagram: (3 ((D) is defined without the notion of an input<br />

subtraction; the feedback network includes reactive elements to provide the phase shift required<br />

-205-


Analog Electronics / Feedback Circuits<br />

for positive feedback; and, of course, there is no external signal source. Nonlinearities that limit<br />

the signal amplitude invariably arise but do not appear in this linear model. The switch helps us<br />

examine the loop gain of the circuit.<br />

Barkhausen Criterion. Assume that A((&) and (3(co) are defined in such a way that no loading<br />

occurs when the switch is closed. The Barkhausen criterion states that there will be sinusoidal<br />

oscillations at frequency C0 o when the switch is closed, provided that with the switch open, the<br />

loop gain is<br />

V f I V t = A{o) 0 )P{co 0 ) = M{co 0 )Z{co 0 ) = 1 (6.43)<br />

When this condition is satisfied, a hypothetical sinusoidal signal generator V; attached to the<br />

input can be removed when the switch is closed, because the amplitude and phase of the signal<br />

fed back to the input are exactly those needed to replace this source. Since the Barkhausen<br />

criterion involves a complex-valued function, it implies two conditions for oscillations, a<br />

magnitude condition and a phase condition. We use the magnitude condition. M(C0 o ) — 1, as a<br />

test to determine whether oscillations can exist in a given circuit. This condition arises from the<br />

physical requirement that the amplifier provide sufficient gain to make up exacdy for energy<br />

losses in the circuit. If Af(C0 o ) exceeds one, the oscillator is self-starting, with the oscillation arising<br />

spontaneously and increasing in amplitude until nonlinearities cause a reduction in M((0 0 )- Some<br />

oscillator circuits require- a signal generator for start-up; however, self-starting circuits are die<br />

norm. Provided that oscillations can occur, the phase condition


Analog Electronics / Feedback Circuits<br />

1/coC<br />

co„+Aco<br />

(c)<br />

Figure 6.19 Wien bridge oscillator: (a) schematic; (b) redrawn to sliow the bridge;<br />

(c) illustration of frequency stability.<br />

To find design equations for this or any oscillator, we apply the Barkhausen criterion. From Fig.<br />

6.19a, P(ffl) is the transfer function<br />

where Z, and Z 2 are the impedances of the parallel RC and series RC circuits, respectively. Since<br />

A(a>) is the gain of the noninverting amplifier, Eq. (6.43) requires<br />

A{G> 0 )P{a> 0 ) =<br />

1 + *2<br />

R \J 1 + j


Analog Electronics / Feedback Circuits<br />

not particularly low values. To give an intuitive idea of what is involved, Fig. 6.19c shows Eq.<br />

(6.46) as the graphical solution of the equation 1/toC = R. The dashed lines show how C0 o<br />

changes with variations in C. (Like the sensitivity expression, this diagram assumes that the<br />

resistor and capacitor pairs in die P circuit are matched.)<br />

The next example demonstrates another useful way to investigate frequency stability of an<br />

oscillator - using SPICE to plot the phase of A(a>)p((o).<br />

Example 6.4 Plot the phase shift of the loop gain for a Wien bridge oscillator with R = 1591.5<br />

Q and C = 1000 pF. Find the change in frequency if the value of the shunt capacitor increases by<br />

10%.<br />

Solution. Figure 6.20a shows the phase-shifting network - Fig. 6.20b the SPICE code. An<br />

asterisk marks the statement that describes C 2 for the second run. Since Eq. (6.46) indicates an<br />

oscillation frequency of 100 kHz, the phase of P(


Analog Electronics / Feedback Circuits<br />

Phase-Shift Oscillator. Figure 6.21a shows the phase-shift oscillator, The phase shifter consists<br />

of three RC sections, The gain element is represented as an ideal inverting amplifier with voltage<br />

gain -K.<br />

To find amplitude and phase conditions from the Barkhausen criterion, we analyze the phaseshifting<br />

circuit of Fig. 6.21b. The transfer characteristic' of such a ladder circuit is most easily<br />

found by systematic analysis from output to input as suggested by the following equations.<br />

' - * •<br />

2<br />

R<br />

l<br />

V: +<br />

1<br />

= 1+ ,VjcoRC)<br />

'<br />

K jaRC) R R<br />

Continuing in this fashion finally gives<br />

2 + jcoRC) R<br />

v i 1<br />

5 1<br />

1 + jcoRC o 2 R 2 C 2+J co 3 R 3 C 3<br />

Thus our design must satisfy<br />

-K<br />

A{O) 0 ){3{CD 0 ) = ^- =<br />

= 1<br />

6 5 1<br />

1 +<br />

jco 0 RC~ CO 2 OR 2 C 2+J o) 3 0R 3 C 3<br />

To find the phase condition, we set the imaginary part of the denominator to zero. This gives an<br />

oscillation frequency of<br />

(6 ' 47)<br />

^-7SE<br />

It is easy to show that the sensitivity to changes in R and C is the same as for the Wien bridge<br />

circuit. Substituting co 0 from Eq. (9,39) into the preceding equation gives the gain condition<br />

-K<br />

TTir 1<br />

Thus the circuit oscillates for any K > 29.<br />

A(


Analog Electronics / Feedback Circuits<br />

Ans. Vj =<br />

jaRC (wRC) 2 V t<br />

Amplitude Limitets. To make an oscillator self-starting and to allow for uncertainties in<br />

parameter values, we usually design the circuit so that the gain condition is exceeded. The<br />

amplitude of the oscillation then increases until some nonlinearity reduces the effective loop gain.<br />

If the signal amplitude becomes too large, the signal traverses a large segment of the nonlinear<br />

transfer characteristic of the active device causing the sinusoidal output waveform to be highly<br />

distorted. There are three basic approaches to controlling the signal amplitude while keeping the<br />

waveform reasonably sinusoidal.<br />

When the gain condition is not gready exceeded, a design can rely on inherent transistor<br />

nonlinearities to limit the signal amplitude to a value corresponding to a reasonably undistorted<br />

sine wave. This does not produce very robust designs, however, because component values are<br />

rather critically related to waveform purity.<br />

The second approach is to insert a special nonlinear component into the loop so that loop gain<br />

begins to diminish with signal amplitude while signal amplitude is still small. A thermistor with<br />

resistance that decreases- by self-heating, or a strategically placed diode are examples of<br />

components used for this purpose. An example is Fig. 6.22a, which shows the phase-shift<br />

oscillator of Fig. 6.21a with two diodes added for limiting. As long as the voltage across resistor<br />

Rj has peak amplitude less than 0.5 V the circuit is approximately the same as Fig. 6.22a;<br />

however, when output amplitude exceeds the forward-biased diode voltage, the diodes introduce<br />

a low resistance in parallel with ly This nonlinear resistance decreases with increasing amplitude,<br />

lowering the gain of the inverting amplifier and thereby limiting the oscillation amplitude. The<br />

following example investigates this limiter in some detail.<br />

Example 6.5 Use PSpice to examine the output waveform of the phase-shift oscillator of Fig.<br />

6.22a without and with the diode limiter. Values are R = 10 kQ, K = 50, and C = 3000 pF. Use a<br />

dependent source with gain 10 5 and input impedance of 1 MQ for the op amp.<br />

Solution. Figure 6.22b is the PSpice code with diodes omitted. The initial condition statement,<br />

".IC V(3) = 0.1," changes the initial voltage at node 3 to a nonequilibrium value to help start<br />

oscillations. Figure 6.22c shows the output without the diode statements. Since the linear circuit<br />

model has no limiting whatsoever, the oscillation amplitude quickly increases to tens of volts in<br />

the simulation. The output of a real opamp would be driven to its saturation limits, giving a<br />

highly distorted output voltage, a statement readily verified using the more realistic operational<br />

amplifier model, e.g. from the EVAL.LIB library.<br />

Figure 6.22d shows the output with the diode statements present in the input code. The<br />

waveform quickly converges to a sinusoidal waveshape, with amplitude limited by the diodes and<br />

virtual ground to the ±0.5 V range.<br />

+<br />

Exercise 6.10 Calculate the expected oscillation frequency in Example 6.5.<br />

Ans. About 1.7kHz with limiting diodes omitted; 2.17 kHz with limiting diodes present.<br />

-210-


Analog Electronics /Feedback Circuits<br />

(c)<br />

(d)<br />

Figure 6.22 Phase-shift oscillator with diode limiter (a) schematicj-(b) PSpice code;<br />

(c) output with limiting diodes omitted; (d) output with limiting diodes included.<br />

For sinusoids of very high quality, a „linear amplifier" can provide limiting with gain that<br />

decreases as amplitude increases. In Fig. 6.23, the voltage-controlled amplifier is the gain element<br />

in a Wien bridge oscillator. A half-wave rectifier with a capacitive load (envelope detector)<br />

produces a dc control signal, V c , proportional to signal amplitude. The follower isolates the<br />

control circuit from the RC bridge so amplitude and phase conditions are unchanged. As<br />

oscillation amplitude increases, V c becomes more negative, and gain automatically decreases. The<br />

detector time constant R D C D should be several periods of the oscillator waveform.<br />

Because it involves nonlinear operation, limiting always introduces "impurity" into the oscillator<br />

output waveform, causing the output to differ from the desired single-frequency sinusoid. One<br />

can describe any periodic waveform as a Fourier series<br />

00<br />

v(0= T,A„cos(na) 0 t+ „)<br />

n=0<br />

For an oscillator, the ideal output is the n = 1 term alone; terms for which n>2 are undesired<br />

impurity terms introduced by the limiter. The smaller these harmonic terms, the greater the<br />

-211-


Analog Electronics / Feedback Circuits<br />

waveform purity. A useful measure of waveform impurity is the percent total harmonic distortion,<br />

THD, where<br />

rms value of harmonic components<br />

THD =<br />

xl00%<br />

rms value of fundamental frequency terms<br />

If we know the fundamental frequency^. = C0 o /27t of the oscillator output, we can use PSpice to<br />

compute THD by adding a SPICE statement ".FOUR". This asks SPICE to compute amplitudes<br />

of the first nine Fourier series components and to use them to estimate THD. Adding the<br />

statement<br />

.FOUR2.0E3V(1)<br />

reveals that the output voltage waveform contains 6.22% distortion (look into the PSpice output<br />

file). The advantage of more sophisticated limiters like Fig. 6.23 is improved waveform purity.<br />

The oscillator designer is often required to meet a THD specification along with specifications of<br />

frequency and amplitude.<br />

Figure 6.23 A limiter that uses sip~-' amplitude to control loop gain<br />

6.7.3 LC Oscillators<br />

Two important oscillators use the thtee-element n structure of Fig. 6.24a for the phase shifter. In<br />

Fig. 6.24b, a transistor biased for small-signal active operation provides the gain. From Fig. 6.24b,<br />

00 (b)<br />

Figure 6.24 Oscillator configuration with three-element ladder phase shifter:<br />

(a) general structure; (b) oscillator with transistor active element<br />

-212-


Analog Electronics / Feedback Circuits<br />

A(a>) = V f = -g m [R 0 \\Z l \\{Z 2 + Z 3 )]<br />

and<br />

z 3<br />

This gives the loop gain<br />

i4(«)fl») =<br />

Z + Zi<br />

~ l s ^<br />

ZJ + Z<br />

„<br />

J + *<br />

A +1<br />

*o Zl<br />

The Barkhausen criterion requires<br />

r<br />

Z 1 Z 2 +Z 1 Z 3 +i 0 (Z3+Z2+Z 1 )<br />

v<br />

'<br />

If each impedance is an LC element, then Z,- = jXj, where X, = fflJL,-for an inductor and X s<br />

//(oC for a capacitor. Then<br />

MZd> = SmKZfo = lzoo M (649)<br />

K<br />

^ -(X l X 2 +X l X 3 ) + jR 0 (X l+ X 2 + X 3 )<br />

><br />

Because die numerator is real, the phase condition is satisfied only if<br />

X x + X 2 + X 3 = 0 (6.50)<br />

When Eq. (6.50) is satisfied, the gain condition in Eq. (6.49) reduces to<br />

-gmV^^ (651)<br />

X 2 +X 3<br />

We next use these conditions to study two specific oscillators.<br />

Colpitts Oscillator. In the Colpitts oscillator, Z 2 is an inductor and Z, and Z, axe capacitors as in<br />

Fig. 6.25. Substituting Z 2 -faL.^ Z, = 1/foC t and Z } = 1//oC } into Eq. (6.50) gives<br />

1 , 1 n<br />

Solving forOf) gives the phase condition<br />

2<br />

C l +C 3<br />

ffl »=^A<br />

(6 - 52)<br />

which establishes the frequency of osculation. A second design condition comes from Eq. (6.51),<br />

If!^=_s^al (6 . 53)<br />

0> o L 3<br />

Substituting C0 o from Eq. (6.52) leads to the second design equation<br />

-213-


Analog Electronics / Feedback Circuits<br />

g m Ro> (6.54)<br />

which specifies the gain required for sustained oscillations.<br />

Exercise 6.11 Compute the sensitivity of the resonant frequency of a Colpitts oscillator to L<br />

Use your answer to estimate the actual resonant frequency of an oscillator designed to operate at<br />

1 MHz if L 2 is 12% high. Ans. -0.5, 940 kHz.<br />

+12V +12V<br />

+12V +12V<br />

Figure 6.26 Colpitts oscillator realizations: (a) circuit diagram for a BJT oscillator;<br />

(b) small-signal equivalent for BJT circuit; (c) MOSFET oscillator<br />

Example 6.6 The Colpitts oscillator of Fig. 6.26a has a transistor biased at 0.5 mA; pV = 120; C c<br />

is a large coupling capacitor. Find values for L^ C, and C, so the circuit oscillates atf, = 1 MHz.<br />

Ignore transistor capacitances.<br />

Solution. Figure 6.26b shows the high-frequency equivalent circuit. At 0.5 mA,&, = 0.02 S and r n<br />

- 6 kQ. The small-signal equivalent circuit is Fig. 6.26b, where<br />

5.5*0 = ^ ||159*Q||112Afi<br />

From Eq. (6.54),<br />

(0.02)(6xl0 3 ) = 120>§-<br />

c l<br />

(6.55)<br />

-214-


Analog Electronics /Feedback Circuits<br />

and from Eq. (6.52),<br />

LoCea = 7TT = 25.3 x 10" 15 (6.56)<br />

1<br />

** (2/rl0 6 ) 2<br />

where C tq = C,C 3 /(C t + Q.<br />

We now have two equations involving C h C s and L. 2 . In a FET design we would make an<br />

arbitrary choice of one component value; however, in a BJT design we must minimize the effect<br />

of r„, since our theory does not include a resistor in this location. Therefore, we select C3 such<br />

that its reactance is much smaller than r x = 5.5 kQ. This gives the third design equation. Using<br />

two orders of magnitude to assure r n has negligible effect gives<br />

1 1<br />

2*10 6 C 3 " 100<br />

orCjS2894pF.<br />

5.5 xlO 3<br />

Since our design depends upon approximation, and since actual component values may not have<br />

exacdy the values we expect, we satisfy Eq. (6.55) by using C,/C f = 75 instead of the limiting<br />

value of 120 to ensure self-starting oscillations. With C } previously selected, this gives<br />

2894<br />

Q= — = 38.6pF<br />

and C, q = 38.1 pF. From Eq. (6.56)<br />

25.3 xlO -15 n, tre TT<br />

L) = 7T = 0.65 5mH<br />

38.6 x 10" 12<br />

The coupling capacitor C c must have reactance oL 2 at 1 MHz. Since the latter is 4.11 kfi, C c<br />

- 10 nF will do.<br />

In Colpitts oscillators, an inductor called an RF choke usually replaces the collector or drain<br />

resistor. The choke coil is a short circuit for biasing purposes but presents an open circuit at RF<br />

(radio frequency) oscillator frequencies. When a choke is used, R, in Fig. 6.25 becomes the<br />

output resistance r„ of the transistor. The RF choke reduces power dissipation of the circuit and<br />

improves the purity of the output waveform.<br />

Exercise 6.12 Figure 6.26c shows a MOSFET oscillator with drain biased at 1 mA through an<br />

RF choke. If transistor parameters are K = 4 mA/V 2 , V, = 1.2 V, and V A = 70 V, find the<br />

condition for oscillation. Ans. 197 > C 3 /C t .<br />

We have seen that SPICE simulations are useful in verifying oscillator designs, examining<br />

distortion, and exploring sensitivity limitations. There are some practical simulation difficulties<br />

when large bypass and coupling capacitors are included in simulations along with components<br />

having short time constants. Sometimes there are convergence problems in the numerical<br />

algoriuims. Sometimes a very large number of oscillator cycles are required before the circuit<br />

settles into steady-state operation. The following example shows a way to avoid time constant<br />

problems in oscillator simulations. An initial dc analysis determines the voltages across the large<br />

capacitors, which we then replace by dc voltage sources of appropriate polarity.<br />

-215-


Analog Electronics / Feedback Circuits<br />

Example 6.7 Use PSpice to show how the oscillations arise in die circuit designed in Example<br />

6.6.<br />

Solution. The SPICE code of Fig. 6.27a uses values determined in die design of Fig. 6.26a.<br />

Following initial dc analysis, capacitors C E and C c were replaced, respectively, by sources VEE<br />

and VCC. Because the oscillation period is expected to be approximately 1 Lis, the initial .TRAN<br />

statement runs the simulation for 30 periods to show die build-up of oscillations.<br />

Figure 6.27b shows that die oscillation is superimposed upon a 9.2 V dc collector voltage. By<br />

measuring die average period of the last nine cycles from the output data, we find that^ = 0.9414<br />

MHz. The oscillations reach steady-state conditions after about 15 Lis; however, longer times<br />

should be expected if die capacitors C E and C c are not replaced by voltage sources.<br />

This simulation demonstrates limiting performed by inherent transistor nonlinearities rather man<br />

by a special limiter circuit Because we used C 3 /C, = 75 in our design, the loop gain is rather high<br />

and die resulting oscillation has peak amplitude of about 3.5 V. Harmonic distortion is about<br />

13%.<br />

Hartley Oscillator. The Hartley oscillator, Fig. 6.28, is Fig. 6.24b with inductors for Z, and Z,<br />

and a capacitor for Z For diis circuit, Eq. (6.50) gives<br />

(0 0 L\~—— + G) o L 3 =0<br />

tf>o c 2<br />

which establishes the oscillator frequency<br />

1<br />


Analog Electronics / Feedback Circuits<br />

-ZmKOok Zm R o C 2<br />

>1<br />

1<br />

1<br />

Substituting Eq. (6.57) into this expression and simplifying gives the gain condition<br />

(6.58)<br />

Figure 6.28 Hartley oscillator small-signal circuit<br />

6.7.4 Quartz Crystal Oscillators<br />

Frequency Stability. We know that frequency stability is an important consideration in an<br />

oscillator. The Colpitts and Hartley design equations show that in these circuits the frequency<br />

depends entirely upon the reactive components in the resonators. We are therefore concerned<br />

with variations in these components due to aging, temperature, and tolerances, especially since<br />

transistor capacitances are sometimes part of the phase shifter.<br />

Reactance<br />

1/(0 C.<br />

Figure 6.29 Frequency stability of Colpitts oscillator<br />

By defining C tq = C,C } /(C, + C,) for the Colpitts oscillator, we can view


Analog Electronics / Feedback Circuits<br />

example, consider a bar of quartz firmly held at the left-hand end but free to flex up and down at<br />

the right-hand end as illustrated in Fig. 6.30. Conducting, e.g. silver, electrodes are plated to the<br />

upper and lower faces of the bar. Under suitable conditions, voltage applied to the electrodes<br />

forces the right-hand end of the bar to move upward. On the other hand, voltage of the opposite<br />

polarity flexes the bar downward. The piezoelectric effect is reciprocal. In other words, if the<br />

terminals are open circuited and a force is applied to flex the bar, a voltage appears across the<br />

electrodes.<br />

\<br />

Figure 6.30 Simplified physical structure of piezoelectric crystal<br />

In quartz crystals, slight; reversible, physical deformations result in an electrical voltage, and,<br />

conversely, applied voltage produces physical deformations. The crystal is thus an<br />

electromechanical device in which electrical excitation and mechanical deformation are tightly<br />

coupled, a feature that makes the crystal highly useful as a transducer. Another unusual property<br />

of a quartz crystal is that, once set in motion, its energy losses per cycle are very slight. Its<br />

electrical equivalent circuit is an LC resonant circuit like Fig. 6.31a, in which the energy losses of<br />

the crystal are embodied in the crystal parameter r, and C is associated with the external holder<br />

that makes electrical contacts to the quartz. Representative component values for a 90 kHz<br />

crystal are L = 137 H, C = 0.0235 pF, r- 15 kQ and C - 3.5 pF.<br />

When used as a frequency-determining element, the crystal is mounted so that it can vibrate<br />

freely at die desired frequency. The mechanical vibrations result in an ac current in the external<br />

circuit. In an oscillator circuit, an amplifier maintains the vibrations. Because quartz is an<br />

extremely stable material, frequency variations due to changes in power-supply voltage or<br />

temperature are very small compared to those of UZ or RC oscillators.<br />

4<br />

C<br />

(a)<br />

(b)<br />

Figure 6.31 Crystal resonator: (a) schematic symbol and electrical equivalent circuit;<br />

(b) approximate reactance curve<br />

We now derive an expression for the reactance of the quartz crystal that allows us to compare the<br />

crystal resonator characteristics with the JJC reactance curves of Fig 6.29. Near resonance where<br />

-218-


Analog Electronics / Feedback Circuits<br />

me crystal is used, r « ©L. With resistance omitted to simplify die development, the impedance<br />

of the crystal in Fig. 6.31a is<br />

jd)L + 1 1<br />

j 5 LCC+ja>(C + C)<br />

jooC jaC<br />

where the second expression follows from multiplying numerator and denominator by<br />

{/


Analog Electronics /Feedback Circuits<br />

The graphical construction of Fig. 6.32b shows why the oscillation frequency of the Pierce<br />

oscillator is virtually independent of variations in resonator capacitance. Another popular crystal<br />

oscillator circuit is obtained by replacing one of the inductors in a Hartley oscillator with a crystal.<br />

Usually, a quartz crystal can vibrate in many different ways called modes. For example, returning<br />

to die bar of quartz fixed at one end, the bar could flex up and down. On the other hand, it could<br />

flex sideways. If the widdi and height of the bar are different, the frequency for the sideways<br />

motion is different from that of vertical flexure. Another possibility is for the bar to twist around<br />

its axis.<br />

End<br />

stationary<br />

A<br />

Stationary<br />

point<br />

(c) Third overtone<br />

Figure 6.33 Overtone vibrations<br />

Commonly, there are overtone vibrations for each mode. For example, several overtone<br />

vibrations are shown for vertical flexure of a bar in Fig. 6.33. The lowest frequency is called<br />

fundamental. The «th overtone frequency is nearly — but not exactly — n times the frequency of<br />

the fundamental vibration. {The amplitudes of vibrations in Fig. 6.33 are exaggerated for clarity.<br />

Actual amplitudes of vibration in quartz crystal are much smaller.)<br />

The flexure modes that are illustrated in Fig. 6.33 are not often used for crystals. (An exception is<br />

32,768-Hz crystal used in electronic watches.) We have discussed this mode mainly because it is<br />

easy to illustrate. Typical high-frequency crystals use shear modes. Crystals are practical as<br />

-220-


Analog Electronics / Feedback Circuits<br />

frequency determining elements for frequencies in die approximate range from 10 kHz to 200<br />

MHz. Below about 30 MHz, the fundamental mode is used. At higher frequencies, overtones are<br />

used.<br />

6.8 Summary<br />

Adding negative feedback to an amplifier reduces sensitivity to parameter variations, increases<br />

bandwidth, and reduces nonlinear distortion. We can also use feedback to increase or decrease<br />

midrange input resistance and (independendy) increase or decrease output resistance. All<br />

improvements involve multiplication or division by the improvement factor 1 +Ap. The closedloop<br />

gain is simultaneously reduced by diis factor; however, diis counts as an improvement once<br />

we realize that die new, lower, gain usually approximates 1/p. By using resistor ratios for P we<br />

can control this closed-loop gain to ±1%.<br />

With four classes of feedback, we can make our nonfeedback amplifier more closely resemble an<br />

ideal voltage, current, transresistance, or transconductance amplifier at midrange frequencies. For<br />

each class of feedback, the definition of gain, A, corresponds to that of the ideal amplifier. The<br />

origin of the feedback signal defines the output circuit of the feedback amplifier-voltage (current)<br />

feedback lowers (raises) output resistance, making the original amplifier more closely resemble a<br />

dependent voltage (current) source. The way the feedback signal subtracts from die input signal<br />

determines die nature of die input circuit. Series (shunt) feedback involves a voltage (current)<br />

subtraction that increases (decreases) input resistance. This makes die amplifier better represent a<br />

voltage- (current-) controlled dependent source.<br />

The resistive P circuits we use in practical feedback amplifiers introduce loading at input and<br />

output. To apply ideal feedback equations to this case, we represent die p circuit by a two-port<br />

equivalent - the one that corresponds to the particular kind of feedback we employ.<br />

To successfully implement feedback or evaluate die designs of others, we must be able to<br />

recognize some standard practical feedback topologies. The difference amplifier structure is a<br />

voltage-series feedback arrangement that bases the input subtraction on an amplifier's differential<br />

input. In classical voltage-feedback circuits, as in the difference amplifier topology, the output<br />

node of die nonfeedback amplifier connects direcdy to the P circuit. Current feedback features<br />

indirect sensing of output current using an unbypassed emitter or source resistor. Classical series<br />

feedback involves a connection from die P circuit to an unbypassed emitter or source resistor in<br />

the input circuit of the nonfeedback amplifier; shunt feedback uses a direct connection from die<br />

P circuit to an input node of the nonfeedback amplifier. The A circuit must provide a 180° phase<br />

shift to facilitate the current subtraction needed for shunt feedback; zero phase shift is required<br />

for series feedback.<br />

To ensure that our feedback amplifier does not oscillate, we examine die magnitude and phase of<br />

its loop-gain function, the product of the A circuit and P circuit gains. If diere exists no<br />

frequency where the loop gain is negative and also greater than one in magnitude, the feedback<br />

amplifier will be stable. Gain and phase margins are two measures of the degree of stability that<br />

might be included in design specifications. If die feedback amplifier is destined to be unstable, we<br />

must compensate it by modifying die open-loop gain curve to produce appropriate gain and<br />

phase margins. This process, called frequency compensation, usually involves adding capacitance<br />

to reduce die bandwiddi of die A circuit.<br />

-221 -


Analog Electronics / Feedback Circuits<br />

Oscillators are circuits intentionally made unstable by positive feedback. Common to all<br />

oscillators are phase shifters, which determine the frequency of oscillation, gain circuits to make<br />

up for energy losses in the phase shifters, and limiters to control oscillation amplitude. The<br />

Barkhausen stability criterion, based upon conditions for positive feedback, gives a complex<br />

valued equation for any oscillator. The Barkhausen conditions give an equation for the frequency<br />

of oscillation and an amplitude condition that must be satisfied for oscillations to occur. In this<br />

chapter only oscillators with sinusoidal output waveforms were considered. Phase-shift and Wien<br />

bridge oscillators use RC circuits for phase shifters; Colpitis and Hartley oscillators employ tuned<br />

circuits.<br />

Important considerations in sinusoidal oscillator design are the harmonic purity of the waveform<br />

and frequency stability, the relative insensitivity of the oscillation frequency to variations in circuit<br />

parameters, and environmental factors. Total harmonic distortion measures the impurity of the<br />

output waveform in terms of the amplitudes of undesired harmonics present. The sensitivity<br />

definitions are useful aids in oscillator analysis and design. SPICE simulations help us address<br />

both distortion and sensitivity issues in a practical fashion. Quartz crystals introduced into<br />

oscillator phase shifters result in great improvements in frequency stability.<br />

Notes<br />

-222-


Analog Electronics / Feedback Circuits<br />

References<br />

1. D. Christiansen (Ed.), Electronics Engineers'Handbook, McGraw-Hill, 1996.<br />

2. A. R. Hambley, Electronics, Macmillan, 1994.<br />

3. J. Keown, PSpice and Circuit Analysis, Maxwell Macmillan, 1991.<br />

4. N. R. Malik, Electronic Circuits, Prentice Hall, 1995.<br />

5. R Mauro, Engineering Electronics, Prentice Hall, 1989.<br />

6. J. Millman and A. Grabel, Microelectronics, 2 nd edition, McGraw-Hill, 1987.<br />

7. J. F. Morris, Introduction to PSpice, Houghton Mifflin, 1991.<br />

8. C. J. Savant, M. £J. Roden and G. L. Carpenter, Electronic Design, The Benjamin/Cummings,<br />

1991.<br />

9. J. Scott, Analog Electronic Design, Prentice Hall, 1991.<br />

10. A. S. Sedra and K. C. Smith, Microelectronic Circuits, Saunders College, 1992.<br />

11. W.J. Tompkins (Ed), Biomedical Digital Signal Processing, Prentice-Hall, 1993.<br />

12. P. W. Tuinenga, SPICE: A Guide to Circuit Simulation and Analysis Using PSpice, Prentice Hall,<br />

1992.<br />

-223-


Analog Electronics /Review Questions<br />

Review Questions<br />

Chapter 1: introduction<br />

1.1 List five examples of electronic systems. Try to think of new examples that have not been<br />

mentioned in this chapter.<br />

1.2 Discuss briefly what the spectrum of a signal is and why it is important.<br />

1.3 Give five examples of useful signals and their spectra.<br />

1.4 List five types of functional blocks found in electronic systems.<br />

1.5 Discuss how analog signals can be converted to digital form.<br />

1.6 List the relative advantages of digital systems compared to analog systems and vice versa.<br />

1.7 Explain why at certain frequencies a physical resistor can be perceived as a capacitor or<br />

inductor.<br />

Chapter 2: Amplifiers<br />

2.1 How does an inverting amplifier differ from a noninverting one<br />

2.2 Draw the voltage amplifier model. Is the gain parameter measured under open-circuit or<br />

short-circuit conditions Repeat for current, transconductance and transresistance models.<br />

2.3 What are "loading effects" in an amplifier circuit<br />

2.4 Draw the cascade connection of two amplifiers. What is the voltage gain of the cascade<br />

connection in terms of the voltage gains of the individual stages<br />

2.5 Define the efficiency of a power amplifier. What form does the dissipated power take<br />

2.6 How is power gain converted to decibels Voltage gain<br />

2.7 Give the input and output resistances of an ideal voltage amplifier. Repeat for other ideal<br />

amplifier types.<br />

2.8 Sketch the gain magnitude of a typical dc-coupled amplifier versus frequency. Repeat for an<br />

ac-coupled amplifier.<br />

2.9 How is a narrowband amplifier different from a wideband one<br />

2.10 Discuss the Miller theorem. Explain its significance to circuit analysis.<br />

2.11 What are the requirements for the gain magnitude and phase of an amplifier so that linear<br />

distortion does not occur<br />

2.12 Sketch the pulse response of an amplifier, showing the rise time, overshoot, ringing and tilt.<br />

Give an approximate relationship between rise time and the upper half-power frequency of a<br />

broadband amplifier. Give an approximate relationship between percentage tilt and the lower<br />

half-power frequency.<br />

2.13 What is the compensated oscilloscope probe What for is it used in electronics<br />

2.14 What is harmonic distortion What causes it Is it a problem for narrowband amplifiers<br />

Explain.<br />

2.15 Discuss briefly intermodulation and crossmodulation.<br />

-224-


Analog Electronics /Review Questions<br />

Chapter 3: Diode Circuits<br />

3.1 Draw the circuit symbol for a diode. Label the anode and the cathode. Make a reference to<br />

die p-type and »-type regions of a corresponding/^ junction.<br />

3.2 Draw the volt-ampere characteristic of a typical diode and label the various regions of<br />

operation.<br />

3.3 What is a Zener diode For what is it typically used What are two other names for it<br />

3.4 Draw a circuit diagram of a simple voltage regulator.<br />

3.5 Draw the volt-ampere characteristic of an ideal 5.8-V Zener diode.<br />

3.6 Write the Schockley equation and define all the terms.<br />

3.7 How does the forward voltage of a silicon diode change with temperature for a fixed value of<br />

current Explain.<br />

3.8 What is an ideal diode Draw its characteristic.<br />

3.9 After solving a circuit with ideal diodes, what check is necessary for diodes initially assumed<br />

to be on Off<br />

3.10 If a nonlinear two-terminal device is modeled by die piecewise-linear approach, what is die<br />

equivalent circuit of die device for each linear segment<br />

3.11 A resistor R a is in series with a voltage source V a . Draw the circuit. Label die voltage across<br />

die combination as v and the current as /. Draw and label the volt-ampere characteristic (I<br />

versus v).<br />

3.12 Draw the circuit diagram of a half-wave rectifier for producing a nearly steady dc voltage<br />

from an ac source. Include a transformer to adjust the voltage level. Draw two different fullwave<br />

circuits.<br />

3.13 What is a clipper circuit Draw an example circuit diagram including component values, an<br />

input waveform, and the corresponding output waveform.<br />

3.14 Repeat Question 3.13 for a clamp circuit.<br />

3.15 Draw die circuit diagram of a two-input diode OR gate. Repeat for an AND gate.<br />

3.16 Logic 1 voltages for a digital system are defined as being larger than 3V. Consider a number<br />

k of two-input OR gates employing silicon diodes. Connect one input of each gate to the<br />

ground (logic 0). Cascade the gates such diat the output of a gate is connected to the<br />

remaining input of another gate that follows in the cascade. Apply the 5-V voltage to die input<br />

and measure die output of the cascade. At what value of k die output voltage can not be<br />

recopiized as logic 1<br />

3.17 Try to construct a logic inverter using diodes. Can you succeed Based on results you<br />

obtained answering Questions 3.16 and 3.17, describe two serious drawbacks of diode logic<br />

circuits.<br />

3.18 Of what does die small-signal equivalent circuit of a diode consist<br />

3.19 How is the dynamic resistance of a nonlinear circuit element determined at a given<br />

operating point<br />

3.20 Sketch die voltage and current waveforms corresponding to large-signal diode switching.<br />

Define and explain the delay times in die diode response.<br />

3.21 What is die Schottky diode What are its properties and applications<br />

3.22 Explain die difference between the photodiode and solar cell.<br />

3.23 What is die construction of an optical isolator List applications of diis device diat are<br />

known to you.<br />

Chapter 4: FET Circuits<br />

4.1 Sketch the simplified physical structure of an »-channel JFET. Label the terminals and die<br />

channel region. Draw the corresponding circuit symbol.<br />

-225-


Analog Electronics /Review Questions<br />

4.2 In normal operation, what bias condition exists between gate and channel of JFET<br />

4.3 Define the pinch-off voltage and I DSS of a JFET.<br />

4.4 Write an equation for the drain current of a JFET in the saturation (pinch-off) in terms of<br />

device voltages.<br />

4.5 Sketch the characteristics of an «-channel JFET. Label the saturation, linear and cutoff<br />

regions. Repeat for the/>-channel device.<br />

4.6 Give the ranges of v GS and v GD in terms of the pinch-off voltage V p for each region (cutoff,<br />

saturation and linear) of an »-channel JFET and depletion MOSFET.<br />

4.7 Sketch the physical structure of an ^-channel depletion MOSFET. Label the terminals and<br />

the channel region. Draw the corresponding circuit symbol. Repeat for a ^-channel<br />

enhancement MOSFET.<br />

4.8 Explain the physical origin of the variations in FET parameters with temperature.<br />

4.9 What is "gate protection" for a MOSFET Why is it necessary<br />

4.10 Draw the load line for a simple FET amplifier.<br />

4.11 Why does nonlinear distortion occur in FET amplifiers<br />

4.12 Draw the diagram of the fixed-bias circuit, the self-bias circuit and the fixed- plus self-bias<br />

circuit for a FET. In general, which circuit maintains the most constant drain current from<br />

device to device Which shows the greatest variation Which is used for enhancement<br />

devices Why<br />

4.13 Draw the small-signal equivalent circuit for the FET including r^<br />

4.14 Give definitions of^n and r rf as partial derivatives.<br />

4.15 Draw the circuit diagram of a resistance-capacitance coupled common-source amplifier.<br />

Repeat for the source follower. Which amplifier would be used if a voltage gain magnitude<br />

larger than unity is needed Which would be used to obtain low output resistance<br />

4.16 Draw the circuit diagram of the FET amplifier most useful if extremely high input resistance<br />

is required.<br />

4.17 What is the function of coupling capacitors With what are they replaced in a midband<br />

small-signal equivalent circuit In general, what effect do the coupling capacitors have on the<br />

gain of the amplifier as a function of frequency<br />

4.18 Compare gain, bandwidth, input and output impedance values of basic FET amplifier<br />

configurations.<br />

4.19 What is the value of ^n for V DS =0 Draw the small-signal equivalent circuit at this bias<br />

point. For what applications is the FET used at this bias point<br />

4.20 Draw the circuit diagram of a CMOS inverter. Repeat for a two-input NOR gate.<br />

4.21 Of what does the input impedance of a CMOS inverter consist<br />

4.22 What is the static power consumption of CMOS gates<br />

4.23 Draw the circuit diagram of the CMOS transmission gate.<br />

4.24 Draw the dynamic circuit model of the »-channel JFET. Repeat for/(-channel device and all<br />

kinds of MOSFET transistors.<br />

4.25 Explain how the channel-length modulation effect is incorporated into the SPICE model of<br />

the FET.<br />

Chapter 5: BJT Circuits<br />

5.1 Draw the circuit symbol for an npn BJT. Label the terminals and the currents. Choose<br />

reference directions that agree with the true current direction for operation in the active<br />

region.<br />

5.2 Repeat Question 5.1 for zpnp transistor.<br />

5.3 In normal operation, which type of bias (forward or reverse) is applied to the emitter-base<br />

junction The collector-base junction<br />

-226-


Analog Electronics /Review Questions<br />

5.4 To forward-bias apn junction, which side of the junction should be connected to the positive<br />

voltage<br />

5.5 Write the Shockley equation for the emitter current of an npn transistor.<br />

5.6 Give the definition of a and p for a BJT. What bias conditions for each junction are assumed<br />

in diese definitions<br />

5.7 Sketch the input characteristic curve for a typical small-signal silicon npn BJT at room<br />

temperature. Sketch the output characteristic curves if P = 100. Isabel the cutoff, active and<br />

saturation regions.<br />

5.8 Draw the load lines on the input and output characteristic planes of the npn BJTfor a simple<br />

amplifier. Repeat for thepnp device.<br />

5.9 Why does distortion occur in BJT amplifiers<br />

5.10 Sketch the output characteristics of a BJT, illustrating the Early voltage and collector<br />

breakdown.<br />

5.11 Explain the base-width modulation effect.<br />

5.12 Discuss the BJT parameter variation with temperature.<br />

5.13 What is the typical extreme variation of P from unit to unit for a given type of BJT<br />

5.14 How does v Bh vary with temperature for a fixed emitter current Assume a small-signal<br />

silicon transistor.<br />

5.15 Draw the large-signal dc circuit model for a silicon npn transistor in the active region at<br />

room temperature. Include the constraints of currents and/or voltages that guarantee<br />

operation in the active region. Repeat for the saturation region. Repeat for the cutoff region.<br />

5.16 Repeat Question 5.12 for &pnp transistor.<br />

5.17 In the active region, how is the base-collector junction biased (forward or reverse) How is<br />

the base-emitter junction biased<br />

5.18 Repeat Question 5.14 for the saturation region.<br />

5.19 Repeat Question 5.14 for the cutoff region.<br />

5.20 Briefly discuss the procedure for dc analysis of a BJT circuit using the large-signal circuit<br />

models.<br />

5.21 Draw the fixed base bias circuit. What is the principal reason that this circuit is unsuitable<br />

for mass production of amplifier circuit<br />

5.22 Draw the four-resistor bias circuit for the BJT. Give the rule-of-thumb design guidelines for<br />

this circuit.<br />

5.23 Whv are coupling capacitors often used to connect the signal source and the load to<br />

amplifier circuits Should coupling capacitors be used if it is necessary to amplify dc signals<br />

Explain.<br />

5.24 Draw the small-signal equivalent circuit for the BJT.<br />

5.25 Give the formula for determination of r x , assuming that P and the Q-point are known.<br />

5.26 Draw the circuit diagram of a common-emitter amplifier circuit that uses the four-resistor<br />

biasing network. Include a signal source and a load resistance.<br />

5.27 Repeat Question 5.23 for an emitter follower. What resistance did you omit from the bias<br />

Why<br />

5.28 For a small-signal midband analysis of an amplifier, with what do we replace the coupling<br />

capacitors Dc voltage sources Dc current sources Very large inductors<br />

5.29 Outline the small-signal analysis procedure to find the output resistance of an amplifier.<br />

5.30 What are three important features in the structure of a BJT for high P (i.e., for the base<br />

current to be small compared to the collector current in the active region of operation) ,<br />

5.31 Sketch the common-emittet input characteristics of an npn transistor. Indicate the effect of<br />

base width modulation.<br />

5.32 Prepare a table showing the bias conditions (forward or reverse) for the collector junction in<br />

each of the four regions of operation of a BJT.<br />

-227-


Analog Electronics /Review Questions<br />

5.33 Draw the common-emitter h-parameter equivalent circuit for the BJT, labeling each<br />

parameter.<br />

5.34 Draw the hybrid-7l model for the BJT. Characterize each element of the circuit.<br />

5.35 Discuss the P dependence of frequency.<br />

4.26 Compare gain, bandwidth, input and output impedance values of basic BJT amplifier<br />

configurations.<br />

5.36 What is the cascode amplifier What are its properties<br />

5.37 Draw the Ebers Moll dynamic model for the BJT. Characterize each element of the model.<br />

5.38 Draw the circuit diagram of an RTL inverter. Sketch a positive input pulse and the<br />

corresponding output voltage. Label the delay time, the rise time, the storage time, and the fall<br />

time.<br />

5.39 For fast switching, do we want a BJT with a thin base region or a thick base region<br />

Explain.<br />

5.40 Draw the circuit diagram of an RTL inverter including a speed up capacitor and a Schottky<br />

clamp diode. Discuss how the Schottky clamp diode improves switching time.<br />

5.41 What is the function of the rt buried layer under the collector region of an npn BJT on an<br />

IC<br />

Chapter 6: Feedback Circuits<br />

6.1 List four benefits that potentially result from the use of negative feedback.<br />

6.2 What problems are associated with positive feedback in amplifiers<br />

6.3 Under what condition is feedback able to reduce nonlinear distortion Draw the feedback<br />

amplifier circuit and derive appropriate expression.<br />

6.4 Define signal-to-noise ratio. Under what condition is feedback able to improve signal-tonoise<br />

ratio Draw the feedback amplifier circuit and derive appropriate expression.<br />

6.5 Define the following terms: voltage feedback, current feedback, series feedback and parallel<br />

feedback.<br />

6.6 Describe a way to test a circuit for the presence of voltage feedback. Draw a schematic<br />

diagram of such a feedback circuit. Repeat for current feedback.<br />

6.7 In a series feedback, we usually consider the input signal to be a voltage. Explain why. In a<br />

parallel feedback we usually consider the input signal to be a current. Explain why.<br />

6.8 Sketch the circuit diagram of the simple class B amplifier that was discussed in this chapter.<br />

What causes crossover distortion in this circuit<br />

6.9 List four types of feedback and give the appropriate amplifier gain parameter for each type.<br />

Also give the units of P for each type.<br />

6.10 What type of negative feedback should be employed to increase input impedance To<br />

reduce input impedance<br />

6.11 What type of negative feedback should be employed to make the amplifier output behave<br />

as a nearly ideal voltage source As a nearly ideal current source<br />

6.12 What type of (negative) feedback should be used to obtain a nearly ideal current amplifier<br />

Transconductance amplifier Voltage amplifier Transresistance amplifier<br />

6.13 Draw the circuit diagram of a negative feedback amplifier, including a resistive feedback<br />

network for series current feedback; for parallel current feedback; for series voltage feedback;<br />

for parallel voltage feedback. In each case give the value of the feedback ratio in terms of the<br />

resistor values. Assume an amplifier having a differential input.<br />

6.14 In series feedback we usually try to select small values for the resistors in the feedback<br />

network. Explain why.<br />

6.15 In parallel feedback we usually try to select large values for the resistors in the feedback<br />

network. Explain why.<br />

-228-


Analog Electronics /Review Questions<br />

6.16 In voltage feedback we usually try to select large values for the feedback resistors. Explain<br />

why.<br />

6.17 In current feedback we usually try to select small values for the feedback resistors. Explain<br />

why.<br />

6.18 Define gain margin and phase margin for a feedback amplifier.<br />

6.19 What are the rule-of-thumb minimum values of gain margin and phase margin used in<br />

design of feedback amplifiers<br />

6.20 Can a single-pole amplifier become unstable if negative feedback having constant feedback<br />

ratio P is employed<br />

6.21 Explain the Barkhausen criterion of oscillations in a feedback circuit.<br />

6.22 Draw the circuit diagram of a Wien-bridge oscillator. Explain its operation.<br />

6.23 Draw the circuit diagram of a Wien-bridge oscillator. Explain its operation.<br />

6.24 Explain the role of amplitude limiters in oscillator circuits. Draw schematic diagrams of<br />

diode and FET limiters circuits used to stabilize the output waveform generated by a Wienbridge<br />

oscillator.<br />

6.25 Draw a block diagram of LC oscillators. Modify it to obtain (a) Collpits, and (b) Hartley<br />

oscillator.<br />

6.26 Briefly describe the piezoelectric effect.<br />

6.27 What is a crystal as the term used in relation to oscillator circuits<br />

6.28 Draw the equivalent circuit of a crystal and sketch its reactance versus frequency. Label the<br />

series resonant frequency and the parallel resonant frequency.<br />

6.29 A crystal has a fundamental mode at 10 MHz. What is the approximate frequency of the<br />

second overtone Third overtone<br />

6.30 Briefly discuss the way to use quartz crystal to stabilize the frequency of a conventional LC<br />

oscillator.<br />

-229-


Analog Electronics /Problems<br />

Problems<br />

Chapter 1: Introduction<br />

1.1 A sinusoidal signal source has an open-circuit voltage of 10 mV and a short-circuit current of<br />

10 u A. What is the source resistance<br />

1.2 Give expressions for the sine-wave voltage signals having:<br />

(a) 10-V peak amplitude and 10-kHz frequency,<br />

(b) 120-V rms and 60-Hz frequency,<br />

(c) 0.2-V peak-to-peak and 1000-rd/s frequency,<br />

(d) 100-mV peak and 1-ms period.<br />

1.3 Illustrate the composition of a square-wave signal by sketching the first four terms of its<br />

Fourier series and then by performing graphical summations.<br />

1.4 For a square-wave audio signal, what fraction of the available signal energy is perceived by an<br />

average adult listener of age 40 whose hearing extends only to 16 kHz<br />

1.5 What fraction of the energy contained in a square wave of frequency / and peak-to-peak<br />

amplitude U is contained in the harmonic at frequency 9ft<br />

Chapter 2: Amplifiers<br />

2.1 A signal source with an open-circuit voltage of V s = 2mV rms and an internal resistance of<br />

50kA is connected to the input terminals of an amplifier having an open-circuit voltage gain of<br />

100, an input resistance of lOOkQ and an output resistance of 4Q. A 4-fi load is connected to the<br />

output terminals. Find the voltage gains A„ = vjv t and A r - vjv t Also find the current gain and<br />

the power gain.<br />

2.2. A certain amplifier has an open-circuit voltage gain of unity, an input resistance of 1MQ, and<br />

an output resistance of 100Q. The signal source has an internal voltage of 5mV rms and an<br />

internal resistance of lOOkfl. The load resistance is 50£X If the signal source is connected to the<br />

amplifier-input terminals and the load is connected to the output terminals, find the voltage<br />

across the load and the power delivered to the load. Next consider connecting the load direcdy<br />

across the signal source without the amplifier and again find the voltage and power. Compare the<br />

results. What do you conclude about the usefulness of a unity-gain amplifier in delivering power<br />

to the load<br />

2.3. An amplifier has an open-circuit voltage gain of 100. With a 10-kQ load connected, die<br />

voltage gain is found to be only 90. Find the output resistance of the amplifier.<br />

2.4. The output voltage v, of the circuit of Fig. P2.4 is lOOmV with the switch closed. With the<br />

switch open, the output voltage is 50mV. Find the input resistance of the amplifier.<br />

4-r"T-4-(><br />

v,(0<br />

—c<br />

Figure P2.4<br />

-230-


Analog Electronics /Problems<br />

2.5. Two amplifiers have the characteristics shown in Table P2.5. If the amplifiers are cascaded in<br />

order A-B, find the input impedance, output impedance and an open-circuit voltage gain of the<br />

cascade. Repeat if die order is B-A.<br />

2.6 A certain amplifier has an input voltage of lOOmV rms, an input resistance of lOOkfl, and<br />

produces an output of lOVrms across an 8-Q resistance. The power supply has a voltage of 15V<br />

and delivers an average current of 2A. Find the power dissipated in die amplifier and the<br />

efficiency of the amplifier.<br />

2.7. An amplifier has an input voltage of lOmV rms and an output voltage of 5V rms across a 10-<br />

Q load. The input current is l^iA rms. Find the input resistance. Find the voltage gain, current<br />

gain and power gain as ratios and in decibels.<br />

2.8. An amplifier has a voltage gain of 30dB and a current gain of 70dB. What is the power gain<br />

in decibels If the input resistance is 100k£2, what is the load resistance<br />

2.9. Find the voltage across a 50-fi resistance corresponding to (a) lOdBV, (b) -30dBV, (c) 10<br />

dBmV, (d) 20dBW.<br />

2.10. Find the power levels in watts corresponding to (a) 20dBm, (b) -60dBW, and (c) lOdBW.<br />

2.11. An amplifier has an input resistance of lOOCt, an output resistance of 10Q, and a shortcircuit<br />

current gain of 500. Draw the current and voltage amplifier models for the amplifier,<br />

including numerical values of all parameters. Repeat for transconductance and transtesistance<br />

models.<br />

2.12. Amplifier A has an input resistance of 1MQ, an output resistance of 200Q and an opencircuit<br />

transresistance gain of 100MQ. Amplifier B has an input resistance of 50fi, an output<br />

resistance of 500k£2 and a short-circuit current gain of 100. Find the voltage amplifier model for<br />

the cascade of A followed by B. Find the corresponding transconductance amplifier model.<br />

2.13 Repeat Problem 2.12 if the order of the cascade is changed to B-A.<br />

2.14. An ideal transconductance amplifier having a short-circuit transconductance gain of 0.1 S is<br />

connected as shown in Fig. P2.14. Find the resistance JR,,= vji x seen at the input terminals.<br />

Figure P2.14<br />

-231-


Analog Electronics /Problems<br />

2.15. Repeat Problem 2.14 if the amplifier has an input resistance of lkft, an output resistance of<br />

20Q and an open-circuit transresistance gain of 1 Okfi.<br />

2.16. An amplifier has an input resistance of 1Q, an output resistance of 1Q, and an open-circuit<br />

voltage gain of 10. Classify this amplifier as an approximate ideal type and find the corresponding<br />

gain parameter. In deciding of an amplifier classification, assume that the source and load<br />

resistances are on the order of lkQ.<br />

2.17. Repeat Problem 2.16 if the input resistance is 1MQ, the output resistance is lMft, and the<br />

open-circuit voltage gain is 100.<br />

2.18. In a certain application, an amplifier is needed to sense an open-circuit voltage of a source<br />

and force current to flow through a load. The source and load resistances are variable. The<br />

current delivered to the load is to be nearly independent of both the source resistance and load<br />

resistance. What type of amplifier is needed If the source resistance varies from 1 to 2kQ and<br />

this causes a 1-% decrease in load current, what is the value of the input resistance If the load<br />

resistance varies from 100 to 300Q and this causes a 1-% decrease in load current, what is the<br />

value of the output resistance<br />

2.19. The input signal to an amplifier is v,(t) = 0.01COS(2000TI/) + 0.02cos(40007t/). The gain of the<br />

amplifier as a function of frequency is given by<br />

100<br />

A ~ \ + j(f/1000)<br />

Find an expression for the output signal of the amplifier as a function of time.<br />

2.20. The input signal to an amplifier is the same as in Problem 2.19. The complex gain of the<br />

amplifier at lOOOHz is 100Z-45 0 . What complex value must the gain have at 2000Hz for<br />

distortionless amplification<br />

2.21. Consider the simple low-pass filter shown in Fig. P.21.<br />

(a) Find the complex gain A = V2I Fir as a function of frequency. What are the magnitudes of A<br />

at dc and at very high frequencies Find the half-power bandwidth B of the circuit in terms of R<br />

and C.<br />

(b) Consider the case for which the capacitor is initially uncharged and v,{l) is a unit step function.<br />

Find v 2 (t) and an expression for its rise time t r in terms of R and C.<br />

(c) Combine the results found in parts (a) and (b) to obtain a relationship between bandwidth and<br />

rise time for the circuit. Compare the results to Equation (2.22).<br />

R<br />

—O<br />

Figure P.21 Low-pass filter<br />

2.22. Consider the simple high-pass filter shown in Fig. P.22a.<br />

(a) Find the complex gain A = V2I Vt as a function of frequency.<br />

(b) What is the magnitude of the gain at dc At very high frequencies Find the half-power<br />

frequency in terms of R and C.<br />

-232-


Analog Electronics /Problems<br />

(c) Consider the input signal shown in Fig. P.22b. Assuming that the capacitor is initially<br />

uncharged, find an expression for the output voltage v^t) for / between 0 and T. Assuming diat<br />

RC is much greater than T, find an approximate expression for percentage tilt<br />

(d) Combine the results of parts (b) and (c) to find relationship between percentage tilt and halfpower<br />

frequency.<br />

v;(0<br />

0 T t<br />

(a)<br />

(b)<br />

Figure P22 High-pass filter (a), its input signal (b)<br />

2.23. An audio amplifier is specified to have half-power frequencies of 15Hz and 15kHz. The<br />

amplifier is to be used to amplify the pulse shown in Fig. P.22b. Estimate the rise time and<br />

percentage tilt of the amplifier output The pulse width T is 2ms.<br />

2.24. The gain magnitudes of several amplifiers are shown versus frequency in Fig. P.24. If the<br />

input to the amplifiers is the pulse shown in the figure, sketch the output of each amplifier versus<br />

time. Give quantitative estimates of as many features of each waveform as you can.<br />

A(/)<br />

(a)<br />

100kHz<br />

Figure P24<br />

2.25. The input signal and the corresponding output are shown for several amplifiers in Fig. P.25.<br />

Sketch the gain magnitude of each amplifier versus frequency. Give quantitative estimates of as<br />

many features on die gain sketches as you can.<br />

2.26. (a) A 1-kHz sinusoid is applied to the input of a nonlinear amplifier. list the frequencies of<br />

at least six frequency components that might be present at the amplifier output.<br />

-233-


Analog Electronics /Problems<br />

(b) Repeat if the input is the sum of a 1-kHz sinusoid and a 1.1-kHz sinusoid.<br />