Introduction to Basic Electronics Concepts
Introduction to Basic Electronics Concepts
3 Introduction to Semiconductors 81
6 IC Applications 187
Index 275
1 Resistors, Capacitors, and Voltage
My approach has been to start with the fundamentals and build up your
knowledge step by step. In chapter 1 you'll begin working with just two of the
simplest electronic components, and will learn how to combine them into
circuits. Then you will add more components and move on to more complicated
circuits.
If you are completely unfamiliar with electricity, you may find that the review
in chapter 8 is not enough. In this case it makes sense to prepare yourself for the
study of electronics by reading Basic Electricity, which is also part of this series
of Self-Teaching Guides.
The basic principles of electronics are ultimately derived from physics. Many
of these principles can be expressed most simply and completely as algebraic
equations. In an equation, letters and symbols are used to represent the
components and variables used in electronics. Algebra provides us with
techniques for stating principles and solving problems using these letters and
symbols. You will find a number of equations used in this book. If it has been
some time since you used algebra, you will find a review of it in appendix II. If
you are completely unfamiliar with algebra, there are many textbooks on
elementary algebra available. An understanding of trigonometry will help in
some sections of the book, but it is not required and you will be able to proceed
without it.
One feature of this book that I think you will find enjoyable and useful is the
Learning Circuits. A Learning Circuit is an electronic circuit you can build
yourself, which will help you to understand a principle of electronics. In order to
build the Learning Circuits you need a supply of basic components. These are
not expensive and are readily available at electrical supply stores. Each Learning
Circuit begins with a list of the components you need to build it, and has
complete instructions and drawings to help you put it together. A complete list of
all the components needed for every Learning Circuit appears in appendix I.
Appendix I also has information about the testing equipment used in the
Learning Circuits, and some general advice about putting together electronic
circuits.
What Is Electronics?
The word electronics has two different, though closely related, meanings. This
can be confusing, but you will find you can quite easily tell which meaning is
intended. In the first definition, electronics is "the study of voltage and current
waveforms that vary in time." When this meaning is intended, the word
electronics is singular, and in a sentence it is used with a singular verb. For
example, one would say, "Electronics is the study of voltage and current
waveforms."
You will be able to tell whether electronics refers to the study or to the
devices by observing the way the word is used in a sentence and the context in
which it is used. In ordinary conversation, the second sense of the word is used
most often. If you walk into almost any large department store, you will find a
section of the store called electronics. It's the section where you can buy DVDs,
CD players, and so on. Clearly, the word refers to electronic devices. In this
book, on the other hand, the word electronics refers most of the time to the study
of voltage and current waveforms. That is why this section is called "What Is
Electronics?" and not "What Are Electronics?"!
Throughout this book you will find experiments in electronics you can do
yourself. They are called Learning Circuits, and they have been designed to give
you a hands-on sense of the way electronic circuits work. A circuit is a group of
interconnected electronic components. They perform such tasks as amplification,
waveform generation, filtering, signal sensing, signal switching, logic, radiation,
and electromagnetic field detection.
Don't worry if you don't know what these functions are. By the time you
finish this book you will be familiar with all of them. They are the functions that
make up radios, VCRs, stereo amplifiers, telephones, and all the other electronic
devices we use. You will not be able to design these electronic products when
you finish this book (that requires more advanced study), but you will have a
much better appreciation of how they work, and you will be well prepared to
take the next step toward learning to design them yourself.
In this chapter there are 8 Learning Circuits. These first experiments will
show you different ways of connecting two basic electrical components: resistors
and capacitors. Before you create your first cir-cult, however, you need some
equipment and some understanding of how to use it.
So, should you dive in and immediately purchase these two pieces of
equipment? That will be your decision, but be sure to read appendix I,
"Preparing to Use the Learning Circuits," first. As you'll see, the equipment is
costly. Before making the purchases, you might want to go through at least a
chapter or two of this book, studying the Learning Circuits and the drawings that
accompany them. You can certainly learn a great deal this way. Then, if you find
you are still excited about electronics, you can look for some used equipment
and start making your own observations.
For the first few Learning Circuits, you can use another piece of equipment
called a multimeter. As the name implies, this is a multiuse measuring device
that can function as a voltmeter, ohmmeter, or ammeter. Multimeters are not
very expensive, and they can measure ac and dc volts, dc current, and resistance.
What they cannot do is show waveforms-for that an oscilloscope is needed.
Circuits also need a source of power, but using utility power from a wall plug
poses a safety hazard. To resolve this difficulty, all of the ac sources in the book
make use of an ac adapter. An adapter is an Underwriter's approved transformer
that supplies a source of low-voltage ac power. (Underwriter's Laboratory is a
testing organization that approves electrical hardware for use by consumers.)
Connecting circuit components together requires some tinned bus wire, some
insulated wire, some solder, and a soldering iron. Simple circuits can be
connected using clip leads or test leads. A test lead is an insulated wire that has
mechanical clips (alligator clips) on each end. In some circuits you can make a
connection simply by twisting leads together. You can also purchase a circuit
board that has a grid of holes so that tie pins can be pressed into the holes, and
you can solder components to these pins. (Just be sure not to cut the leads short
until you are certain they are resting in their final location.) But when circuits
become more complicated, soldering works best. You will find a stepby-step
description of soldering in appendix 1.
You will also need some basic tools, which you may already own: a pair of
needle-nosed pliers and wire cutters.
Finally, you will need a workstation to do the Learning Circuits, which does
not need to be more than a few feet of counter space near a wall outlet. Ideally
this should be a place where you can leave your equipment out and available
while you experiment with the various Learning Circuits.
Each of these waveforms has its own particular use. Sine waves are used to
test the response of circuits. A sine wave is sinusoidal in character. A sine wave
is often referred to as a sinusoid. See Figure 1.1. This waveform is used because
the currents and voltages in a linear circuit are all sine waves. Square waves are
valuable because they provide information about circuit behavior not easily seen
with sine waves. A square wave voltage can transition symmetrically around 0
volts or transition from 0 to a peak voltage level once per cycle. The transition
time or rise time should be short compared to the time of one cycle. This makes
it difficult to generate a 2-MHz square wave, as the transition times should be
around 5.0 ns. Shorter transition times raise the cost. Triangular waves are useful
because the voltage slopes are constant. However, triangular waves are not
generally used in testing. These three waveforms are shown in Figure 1.1. For
more on these concepts, see chapter 8.
An output cable is usually supplied with the waveform generator (which can
also be called a signal generator or a function generator). The cable can have
alligator clips on the end so that it can be connected to various points in a circuit.
The outer conductor of the cable is called the zero reference conductor of the
signal. It is also called a shield, a ground, or the common conductor. (The word
ground is often used to mean "earth" but this is not always the intended
meaning.) It connects to the ground or common of the circuit you are testing. In
some generators this shield is connected to the safety or green wire of the power
conductor. For most testing, you should remove this connection link or strap.
The circuit you are testing may already be connected to ground. If this is the
case, then two connections to ground can be troublesome. This can be checked
by measuring a low resistance from the common output lead to the third pin on
the power cord.
If the voltage probe is connected to a 1-kHz sine wave voltage, a single sine
wave will be displayed on the screen. If the sine wave frequency is 2 kHz, then
two full sine waves will be displayed. When the dot makes many sweeps per
second, the screen pattern appears stationary.
You can observe the oscilloscope display in slow motion by observing a 1-Hz
sine wave from a function generator, with the sweep frequency on the
oscilloscope set to 1 Hz. At this slow rate you will be able to see the dot move
across the screen, writing a sine wave pattern over and over. A sine wave voltage
display is shown in Figure 1.2.
It is worthwhile spending a little time with the oscilloscope and the function
generator before you get started on the Learning Circuits. I can't tell you exactly
how to work the controls, as there are many different designs. You will have to
hook up the oscilloscope to the function generator and play with the dials until it
becomes clear. Don't worry, you can't hurt yourself or the equipment, so go
ahead and experiment.
Figure 1.2 A sine wave displayed on an oscilloscope
An oscilloscope has one or two input probes. The probe tip is designed to
connect to points in a circuit. The grounding clip on the probe is usually
connected to the zero reference or ground of the circuit. At its other end (inside
the oscilloscope) the grounding clip connects to the oscilloscope frame and to
the power safety or green wire (the third plug in a three-pronged electrical plug).
This connection to ground is required by the National Electrical Code.
Voltages
The symbol used for a voltage source is either the letter V in a circle or the
symbol for a battery. A lowercase v refers to a changing voltage. The polarity
(plus or minus) of a dc voltage will always be indicated.
We will assume that the voltage source can supply the current demanded by
the circuit without changing voltage. This is referred to as an ideal voltage
source. In actuality voltage does change with load, but assuming an ideal source
simplifies the discussion.
If the voltage is a step function or square wave, it will be clearly stated in the
text.
We are now ready to begin using our first electronic components, resistors and
capacitors. These two components are found in most electronic equipment
because they do very basic and important jobs needed in every circuit.
Resistors are the most common electrical component (see chapter 8). They are
used to limit the flow of current in a circuit. By comparison a conductor offers
very little opposition to current flow. There are many types and sizes of resistors.
In electronics, resistors are apt to be small cylinders that are about a half-inch
long. This is the circuit symbol for a resistor:
Capacitors are the second most common component. Their basic function is
to store electrical field energy. This field energy requires electric charge on the
plates of the capacitor. (See chapter 8 for discussion of capacitors and electric
charge.) The ratio of voltage to charge is called capacitance. Since it takes time
to store energy, capacitors can be used to control frequency response, provide
filtering action, provide timing, and store energy in power supplies. Capacitors
are found in almost every circuit design. The circuit symbol for a capacitor is
In the next sections we will be examining the way resistors and capacitors
respond to various voltage waveforms. You will recall that a waveform
generator produces sine waves, square waves, and triangle waves. Sine wave
voltages are the only waveform that keep the same shape in any combination of
resistors and capacitors.
One way to study resistors and capacitors is to apply a "step function" to the
circuit. A step function is a voltage that changes from one value to another. In
many cases a low-frequency square wave can be used as a step function. Digital
circuits make extensive use of square waves and step functions.
In practice, series batteries should be of the same type so that the batteries
will last the same length of time.
Figure 1.3 Batteries in series
LEARNING CIRCUIT 1
2 9-V batteries
1. Set the multimeter to "volts" and to a scale appropriate (the voltage
can be easily read on that range) for measuring 9 to 18 V (usually 25
V). Practice using the multimeter (which is now functioning as a
voltmeter) by measuring the voltage of a battery. It should read 9 V.
3. Tie the common voltmeter lead to the connection between the two
batteries. Note that the ends of the batteries are at +9 V and -9 V.
When batteries are placed in series, any one of the battery terminals can be
called 0 V. If the midpoint or jumper between the two series 9-V batteries is
called 0 V, then the other two terminals are at -9 V and +9 V. The potential
difference between the outer terminals is 18 V.
Batteries may be placed in parallel provided their voltages are equal. Connect
the plus terminals together and connect the minus terminals together. The
resulting voltage is the same as one of the batteries, but the current capability is
increased. A problem with this arrangement is that if one of the batteries
becomes weak, the strong battery will drain into the weak battery.
Consider a 300-52 resistor in series with a 2-ku resistor. The 200 S2 must be
expressed as kS2 or the 2 kS2 must be expressed as ohms. The answer is 2 kS2 +
0.3 kL = 2.3 W. The other solution is 2,000 S2 + 300 0 = 2,300 Q.
where RT is the total resistance and R,, R2, and R3 are the resistances of the
three parallel resistors. This idea can be extended to any number of resistors.
LEARNING CIRCUIT 2
3 1,000-Q resistors
The current in the loop is given by Ohm's law. By convention, the direction of
the current is out of the positive terminal of the battery. The current level for a 6-
V battery and a 1,000-0 resistor is 6 mA. The power dissipated in the resistor is
given by V z/R. (See chapter 8 for discussion of these concepts.) To use this
equation, the voltage must be expressed in volts and the resistor in ohms. The
power is 36/1,000 W or 36 mW. If the voltage were 6 V ac, the answer would be
the same.
Standard carbon resistors are commercially available that cover the range 10
Q to 22 MS2. These resistors are available in 'a-W, %-W, 1-W, and 2-W sizes. It
is good practice to avoid using resistors at more than one-half their wattage
rating. In most circuit applications it is convenient to use resistors of one wattage
size, as X-W size is a typical power level. This means many resistors are rated
higher than they need to be. The standard resistor values, with accuracy to within
20% of the stated value, in the range from 10 to 100 92 are 10, 12, 15, 18, 22,
27, 33, 39, 47, 56, 68, and 82 Q. These same multiples are available in every
decade. For example, resistors are available at 22, 220, 2.2 k, 22 k, 220 k, 2.2 M
and 22 M. Resistors with accuracy to within 10% and 5% of their stated value
are also available.
On the other end of the resistor, 10 different colors plus gold and silver are
used, in three bands. The first two bands indicate numeric values and the third
band indicates the number of added zeros. The numeric meaning of each color is
listed in the following table. Note that gold and silver on this end have different
meanings than they have on the other end.
A resistor with bands of brown, black, and brown reads 100 Q. The first
brown is 1, the black band is 0, and the second brown indicates one added 0. The
resistor with bands of brown, green, and green indicates 1,500,000 Q. The brown
is 1, the first green is 5, and the last green says that you add five zeros. Bands of
green, blue, and gold indicate 5.6 ohms. Green is 5, blue is 6, and the gold says
you divide by 10.
If you have any doubt as to the value of a resistor, use your ohmmeter to
verify the value.
Note that when you use your fingers to clamp the leads of the ohmmeter to
the resistors, your fingers become part of the circuit. The resistance between
your fingers can be as low as 10,000 Q. The resistance depends on the
individual, the surface area, the finger pressure, and the moisture and oils in the
skin. For resistors over 1,000 Q hold the ohmmeter by the plastic handles, and
keep your fingers out of the circuit. When measuring a resistor, your fingers
should not provide a parallel path for current flow, or the answer will be
incorrect.
The current that flows in an ideal resistor depends only on the instantaneous
voltage. If the voltage is a sine wave, a step function, or a square wave, the
current waveform is exactly the same. When the resistor value is very large or
very small, there are exceptions to this statement. Resistors with low resistance
values are influenced by their series inductance, and resistors with high
resistance values are influenced by a shunt parasitic capacitance. These effects
are important at high frequencies. For the circuits we will discuss in this book,
these effects can be ignored.
Two resistors in series across a voltage source form a voltage divider. This
voltage divider circuit is shown in Figure 1.9.
One side of the voltage source is usually called the reference conductor or
zero of potential. The voltage at the junction between the two resistors is a
fraction of the total voltage. The purpose of the voltage divider is to produce this
reduced potential. The ratio of voltage drops is equal to the ratio of resistance
values. The sum of the two voltage drops is equal to the source voltage. In
Figure 1.9 the negative side of the voltage source is the reference conductor. If
the resistors are equal, the attenuation factor is 2. The current in this circuit is the
voltage divided by the sum of the resistors, or I = V/(R1 + R2). The voltage
across the first resistor is this resistance times the current, or V, = I X R, = V x
R,/ (R, + R2). The voltage across the second resistor is given by V2 = V X
R2/(R1 + R2). For example, if Rl = 2 kS2 and R2 = 8 W, their sum is 10 kS2. If
the voltage is 10 V, the current is 1 mA. The voltage across the 2-kc2 resistor is I
X R2, or 2 V. If the negative terminal of the battery is at 0 V, then the
connection between the two resistors is +2 V. If the positive terminal of the
battery is at 0 V, then the junction between the resistors is at -8 V
Figure 1.9 A voltage divider
For sine wave voltages or dc voltages, the power dissipated in each resistor
can be calculated three different ways (see chapter 8). The voltage times the
current, or V x I, is the simplest way. The other two ways are I2R and V2/R. The
units must be in ohms, amperes, and volts to get an answer in watts. Using the
equation P = V x I, the power dissipated in R1 is 2x 0.001=2mW.
LEARNING CIRCUIT 3
3 1,000-0 resistors
1 9-V battery
Source Resistance at DC
When current is taken from a practical voltage source, the voltage drops. This is
the result of current flowing in an internal resistance. An internal resistor and an
external load resistor form a voltage divider (attenuator). The circuit is shown is
Figure 1.10.
Internal resistance of any circuit at the point of interest is the ratio of voltage
change to current change.
One approach to determining the internal resistance (also called the internal
impedance) of a circuit is to add a load resistor and note the voltage change. The
voltage before the load is applied is called the open circuit voltage. The change
in voltage, or voltage difference, is the open circuit voltage minus the voltage
when current is flowing. The change in current is the loaded voltage divided by
the load resistance. The ratio of these two numbers is the internal resistance.
For example, assume the open circuit utility power is 120 V at 60 Hz. A 12-
52 load resistor drops the voltage to 119 V. The voltage difference is 1.0 V. The
current is 119/12 = 9.917 A. The internal resistance is 1/9.917 = 0.1008 Q. The
internal resistance of an ideal voltage source is zero. (Warning: Do not attempt
to make this measurement on an open power connection. There is a good chance
of electrical shock.)
Figure 1.10 The internal resistance of a dc voltage source (also called
source resistance)
The voltage divider in Figure 1.9 has an equivalent source resistance (series
resistance) that can be determined by the ratio of open circuit voltage to short
circuit current. This is known as Thevenin's theorem. The open circuit voltage is
Vo = VR,/(R1 + R2). The short circuit current is Is(: = V/R2. The ratio of Vo/Isc
= R1R2/(R1 + R2). This is the resistance of the two resistors in parallel, or the
source resistance. In order to lower the source resistance, more current must flow
in the divider. If both resistors are one-half of their first value, the source
resistance would be one-half of its previous value. For example, consider a 20-V
source where R, and R2 are both = 10 M. The equivalent circuit is a 10-V source
in series with 5 W. If R, and R2 are each 5 kS2, the source impedance is 2.5 kQ.
When a voltage is applied across two parallel resistors as in Figure 1.11, the
current in each resistor follows Ohm's law.
The total current can be calculated two ways. The first way is find the current
in each resistor and then add the values together. The second way is to compute
the parallel resistance and then calculate the current. If the resistors are 2.5 k1
and 5 k12 and the voltage is 25 V, the currents are 10 mA and 5 mA. The total
current is 15 mA. Using the second method, the resistors in parallel are 1.66 W.
The current is 25/1,666 = 0.015 A or 15 mA.
LEARNING CIRCUIT 4
Determining the Internal Impedance of a Battery
3 1,000-Q resistors
1 9-V battery
Switches
Capacitors
When capacitors are placed in parallel, the capacitances add. Capacitances must
be expressed in the same units for addition. For example, a 0.001 µF capacitor is
in parallel with 500 pF. The problem requires that the 500 pF capacitance be
expressed in µF. This is 0.0005 µF. The total capacitance is 0.0015 µF.
Capacitors in series add like parallel resistors. The sum of the reciprocal
capacitances is equal to the total reciprocal capacitance. For example, consider a
0.2 µF in series with 0.4 µF. 1/0.2 + 1/0.4 = 3/0.4. The total capacitance is the
reciprocal, or 0.4/3 = 0.133 tE Again, to do a calculation all capacitances must
use the same unit.
When a steady current (dc) flows into a capacitor, the voltage rises linearly.
As an example, if a dc current of 1.0 mA flows into a capacitor for 1 ms, the
charge is I x t (see chapter 8, Equation 8.4). Q = 0.001 x 0.001 = 10' coulombs.
This much charge on 0.1 µF is a voltage V = Q/C = 10-6/10-7 = 10 V. The
voltage rises linearly from 0 to 10 V in 1 ms.
The square wave voltage is set to 10 V peak. A 100-ku resistor connects the
square wave voltage to a 0.1 p.F capacitor. If the voltage on the capacitor never
rises to more than 0.1 V, the current is essentially constant at 0.1 mA. The
maximum charge Q is C V = 10-7 x 0.01 = 10-9 C. This charge equals I x t
where I is 10' A. Solving for t yields 10-5 seconds. The square wave must stay
positive for 10-5 seconds and return to 0 for another 10-5 seconds. This is a
frequency of 50 kHz. The waveform across the capacitor will be a triangle wave,
a voltage that rises and falls in a linear manner. This voltage can easily be seen
on the screen of the oscilloscope.
LEARNING CIRCUIT 5
2 0.1-µF capacitors
LEARNING CIRCUIT 6
Observing the RC Time Constant
1 9-V battery
1 1-Me resistor
1 1-µF capicitor
1 SPDT switch
To show that RC has units of time, we can use the following definitions. R =
VII = volts/amperes and C = Q/ V = coulombs/volts. But coulombs = amperes x
time. Therefore RC = (volts/amperes) x (amperes/volts) x time. As you can see,
the volts and amperes cancel, leaving the unit of time.
Figure 1.15 The construction of the circuit in Figure
1.14
The idea of a time constant can be applied a second or even a third time to the
same circuit. In the previous example the voltage falls to 37% in 1 time constant
(10 seconds). The voltage will fall to 37% of 37% in 2 time constants (20
seconds). This is a value of about 13%. In 3 time constants or 30 seconds the
value is about 5%.
Each pointer makes a counterclockwise rotation once per cycle. The height of
the pointers above the horizontal axis represents the instantaneous voltages. The
length of the pointer is the peak value of voltage. When the voltage across the
resistor is maximum, the voltage across the capacitor is 0. When the pointer for
the resistor voltage points straight up, the pointer for the capacitor voltage points
to the right.
Figure 1.16 The rotating pointers for a series RC circuit
This ratio is the impedance of a series resistor and capacitor. If the reactance Xc
of the capacitor is 300 S2 and the resistor R = 400 92, the series impedance is
500 Q.
The angles between the various pointers in Figure 1.16 are called phase
angles. Sine waves that peak at different times are shifted in phase. To discuss
phase relationships, the voltages must be sine waves at the same frequency. The
current in a capacitor is always shifted 90° from the voltage across the capacitor.
The current is said to lead the voltage. The voltage across a resistor is always in
phase with the current. There is no phase shift in a resistor.
The amplitude response has been normalized so that the -3-dB point is at a
frequency of 1 Hz. To find a resistor and capacitor value for a specific cutoff
frequency f, first select a resistor value like 10 ku. If the cutoff frequency is 15
kHz (see chapter 8), then we can find C from the equation Xc = R. In this
example 1/6.28fC = R = 10,000. Solving for C, we obtain C = 1/(6.28 x 15,000 x
10,000) = 0.0011 .tE The response curve in Figure 1.17 is still applicable.
Simply multiply the horizontal scale by 15 kHz.
Figure 1.18 The amplitude and phase response of a first-order low-pass RC
filter
The phase shift through this normalized RC filter starts off near 0 degrees at
frequencies well below 1 Hz. This is because the reactance of the capacitor is
much higher than the resistor and the capacitor performs like an open circuit.
Well above the cutoff frequency, the capacitor acts like a short circuit and the
current is limited by the resistor. This means that the current is nearly in phase
with the source voltage. The voltage across the capacitor lags this current by 90°.
At frequencies well above the cutoff frequency, the phase angle between the
input and the output voltages approaches 90° lagging. At the cutoff frequency,
the phase shift is 45°. The phase shift for this first-order filter is shown in Figure
1.18.
There is a close relationship between phase shift and attenuation slope. In this
RC filter, the amplitude falls off proportional to frequency above the cutoff
frequency. If two RC circuits were to contribute to the attenuation, the amplitude
would fall off proportional to frequency squared. In this situation the phase shift
would double. In general, phase shift is closely related (proportional) to the
attenuation slope.
LEARNING CIRCUIT 7
1 0.001-µF capacitor
1 100-ku resistor
When the roles of the resistor and capacitor are reversed in Figure 1.17, the
output voltage is sensed across the resistor. The frequency and phase response
are mirror images of the low-pass filter. This filter attenuates low frequencies
and passes high frequencies. The circuit and the amplitude and phase response
are shown in Figure 1.19.
The voltage across the resistor plus the voltage across the capacitor must
equal the square wave voltage. In other words, the two top curves in Figure 1.20
add up to a square wave. The leading edge of the square wave comes through
immediately in a high-pass filter. The output voltage falls as the capacitor
charges up. After the leading edge is coupled, the voltage waveform follows an
exponential curve. The voltage drops to 37% of initial value in 1 time constant
equal to the product RC.
A high-pass filter can be used to block an average offset voltage. The filter
allows changing voltages to pass. An application of this filter might be to reduce
the bass or low-frequency response in an audio amplifier. It might be used to
pass a high-frequency carrier signal and reject an audio signal.
LEARNING CIRCUIT 8
SELF-TEST
1. The terminals of the six individual cells of a 12-V battery are exposed. All the
cells are in series. If the negative terminal is labeled 0 V, what are the voltages
at the other cell terminals? What happens if the positive terminal is labeled 0
V?
2. Two 9-V batteries are placed in series. If the connecting point is 0 V, what are
the other two voltages?
5. Two resistors are in parallel. Their resistances are 1 kS2 and 10 kil. What is
the total resistance?
6. Two resistors 510 kS2 and 1.2 MS2 are in parallel. What is the parallel
resistance in units of kits and Ms? (Hint: First rewrite the resistor values using
the same units.)
7. A 10-kc2 resistor measures 5% high or 10.5 kS2. Show that a parallel resistor
of 220 kS2 will reduce the resistor value to near 10 kS2. What is the remaining
error expressed as a percentage of 10 kc ?
10. 100 V ac is placed across a 2-kc resistor. What is the current? What is the
power dissipation?
11. 10 V is placed across a resistor of 0.1 Q. What is the current? What is the
power dissipation?
12. 500 S2 and 1,000 0 are in series across a 15-V battery. What is the current in
the resistors?
13. In problem 12 the 500-52 resistor connects to the negative terminal of the
battery. If the negative terminal of the battery is at 0 V, what is the voltage at
the junction between the two resistors?
14. In problem 13, if the positive terminal of the battery is at 0 V, what is the
voltage at the junction between the two resistors?
16. A 12-V battery is loaded with a 1.2-52 load resistor. The voltage drops to
11.92 V. What is the source resistance?
17. A 15-V dc power supply has a voltage divider consisting of three 2-kS2
resistors. What are the voltages at the two junctions? What is the source
resistance at these two points?
18. Use the circuit of Figure 1.9. If R, is 300 S2 and R2 is 100 S2 and the
voltage is 10 V, what is the voltage at the junction of the two resistors?
19. The voltage in Figure 1.9 is 20 V. If the attenuated voltage is 6 V, what are
the two resistor values if the current drawn is 1 mA? What are the resistor
values if the current drawn is 5 mA?
20. A 9-V battery drops to 8.8 V when a 100-mA load is applied. What is the
internal resistance?
21. A 12-V battery supplies three lamps in parallel. The lamps have resistances
of 2 S2, 3 S2, and 5 Q. What is the total current? Note: The filaments of a
lamp are made from tungsten. The cold resistance of the filament is much
lower than the hot resistance. This problem assumes that the given resistance
values occur when the lamps are illuminated.
22. What is the total capacitance when 0.01 gF is paralleled with 0.1 µF?
23. What is the total capacitance when 330 pF is paralleled with 0.002 µF?
24. What is the total capacitance when 0.001 gF is in series with 500 pF?
28. What is the time constant when the resistor is 100 ko and the capacitor is
0.01 µF? How long are 2 time constants?
29. What resistor forms a 0.1-second time constant with a 0.05 gF capacitor?
30. An RC circuit reaches 95% of final value in 0.1 second. What is the RC time
constant?
34. A high-pass filter has a cutoff frequency of 10 kHz. The R is 100 W. What is
the capacitor value?
35. A high-pass filter is formed using 0.01 gF and 150 kL2. A step voltage of 10
V is applied. How much time elapses before the voltage drops to 3.7 V?
36. In problem 35, how much time is required before the voltage drops to 0.5 V?
37. A high-pass filter has a cutoff frequency of 10 Hz. If the resistor is doubled
in value, what change in capacitor value must be made to maintain the same
cutoff frequency?
ANSWERS
3. 18 V.
4. 7,410 i2.
5. 909 Q.
7. The resistor is corrected to 9.976 kit. The error is 0.024 kit or 0.24%.
8. The current is 2 A. The power is 40 W. The current flows from plus to minus.
12. 10 mA.
13. 5 V.
14. -10 V.
15. -10 V.
16. The voltage changed 0.08 V. The current changed 9.93 A. The source
impedance is 0.008 Q.
17. The voltages are 5 V and 10 V. The source impedance in both cases is 2 kit
in parallel with 4 kit. This is 1.33 W.
18. 2.5 V.
20.20.
21. 1.2 A.
32. The ratio of frequencies is 50:1. The attenuation factor is approximately 50.
37. The reactance must also double. This means the capacitance is half the value.
Objectives
• the way two more electrical components, the inductor and the transformer,
work
The next two electronic components we will be studying are the inductor and
the transformer. Building the Learning Circuits in this chapter will give you a
chance to observe the way these components act in various combinations.
Inductors and transformers have in common the fact that both of them use the
magnetic field. If a review of the magnetic field would be helpful, now would be
a good time to read "The Inductor and the Magnetic Field" in chapter 8.
Induced Voltages
Consider the magnetic field in Figure 2.1. When a test coil of wire is moved in
this magnetic field, a voltage appears on the ends of the coil. If the direction of
motion is reversed, the voltage reverses its polarity. If the test coil is moved
faster, the voltage increases. If the test coil is stationary and the current in the
coil increases, there is a voltage on the test coil. If the current decreases, the
voltage polarity reverses. The voltage is proportional to the rate at which lines of
flux thread the test coil and to the number of turns. If the lines increase at 1,000
lines per millisecond, the voltage might be 10 V. If the lines decrease at 1,000
lines per millisecond, the voltage would then be -10 V. These are induced
voltages.
Lines of flux thread the very coil that carries the current. When the current
changes in the coil of Figure 2.1, there is a voltage at the terminals. This voltage
is in the direction to oppose the change in current. The voltage is proportional to
how fast the current is increasing. This fact is known as Lenz's law. If there is a
changing current, there must be an induced voltage. This is known as Faraday's
law.
Figure 2.2 shows this rise in current for a fixed voltage across an inductor.
A 1-H inductor is not used very often in electronics. Typical inductor values
range from a few microhenries to perhaps 100 millihenries. The millihenry is
0.001 H, abbreviated mH, and the microhenry is 0.000001 H, abbreviated g H.
The maximum rate of change of current for a sine wave current occurs when
the current is 0. In an inductor, when the sinusoidal current is 0, the voltage is
maximum. The rotating pointer system in Figure 2.4 shows the timing
relationship between current and voltage.
LEARNING CIRCUIT 9
1 10-mH inductor
1 10-a resistor
To put together the circuit shown in Figure 2.3, use the construction
layout at the bottom of the figure. When you use a 10-mH inductor, a
voltage of 1.0 V dc across the inductor will cause a current to rise at
100 A/sec. If we limit the maximum current to 0.01 A, we cannot leave
the voltage connected for more than 100 µs. We can do this with a
square wave voltage that stays positive for 100 µs and reverses polarity
for another 100 µs. This is a square wave at 5 kHz. A low-valued
resistor in series with the inductor can be used to measure the current.
The peak voltage across a 10-Q resistor will be 100 mV. Use the
oscilloscope to observe that this voltage is a triangle wave.
In an inductor the voltage leads the current by 90° at all frequencies. Compare
this with a capacitor, where the voltage lags the current by 90° at all frequencies.
Inductors in Series and Parallel
Inductors in series add. All of the units must agree. For example, to add 100
µH to 2 mH requires the 100 µH be converted to 0.1 mH. The sum is 2.1 mH.
Inductors in parallel are treated the same as parallel resistors. The reciprocal of
reactance is susceptance. The susceptances are first added together. The
reciprocal of the total susceptance is the parallel inductance. Consider 3 mH in
parallel with 2 mH. The reciprocals are } ' and ~. The sum is The final answer is
f = 1.2 mH. Again, all the units must agree.
Figure 2.3 The rise and fall of current in an inductor
Figure 2.4 The rotating pointers showing current and voltages in an
inductor
Figure 2.5 The rise in current for step function applied to series RL circuit
At the moment the voltage is applied, the current in the circuit is 0. The entire
voltage appears across the inductor. Equation 2.1 requires that the current start to
increase. This current results in a voltage drop across the resistor. This reduces
the voltage across the inductor, which in turn reduces the rate at which current is
rising. The current continues to increase until it reaches a limiting value
determined by the voltage and the resistor.
This rise in current follows an exponential curve. This is the same as the
voltage curve shown in Figure 1.14 for the RC time constant. The current
reaches 63% of final value in a time given by the ratio L/R. To show that L/R
has units of time, we can use Equation 2.1, which states that volts = (inductance
x amperes)/time. This means inductance = (volts/amperes) x time. Ohm's law
says that resistance = volts/amperes. Dividing inductance by resistance, the units
volts/ amperes cancel, leaving the unit of time. For L/R to equal time in seconds,
the inductance must be in henries and the resistance in ohms.
The voltage across the resistor R in Figure 2.5 is a direct measure of the
current. If a square wave voltage is used instead of the step function, the voltage
waveform across the resistor is a filtered version of the square wave. This
waveform is shown in Figure 2.6.
Impedance is a sinusoidal concept. Our objective is to find the ratio of sine wave
voltage to sine wave current in a series RL circuit. If the current is I, the voltage
across the resistor peaks 90 electrical degrees after the voltage across the
inductor. The rotating pointer system in Figure 2.7 shows these two voltages as
sides of a rectangle.
The circuit in Figure 2.3 is a low-pass filter. When a sine wave of voltage VIN is
applied to the circuit, the voltage across the resistor is the output of the filter
VOUT. The ratio VOUT/ VIN is the gain of the filter. In this filter the gain is
always less than 1.
Small inductors are often placed in circuit leads to restrict the flow of current
at high frequencies. This type of inductor can be as simple as threading a
conductor through a small magnetic core (ferrite bead). The filtering action
depends on the presence of a shunting impedance. If this impedance is not
present, the filter might not function.
Later when we discuss resonant frequency you will see that inductors are not
perfect. Above a certain frequency an inductor functions like a capacitor and the
expected filtering action does not take place. All components have their
limitations, but inductors have several weaknesses. Experience tells a designer
when and how he can use an inductor. Inductors in the microhenry range work
well above 1 MHz when the impedances are below a few hundred ohms.
Figure 2.9 shows a series RLC circuit connected to a sinusoidal voltage. The
impedance of this circuit can be determined by assuming a sinusoidal current
flow and solving for the voltage. The ratio of voltage to current is the
impedance.
The voltage across the inductor is IXL, the voltage across the capacitor is
IXc, and the voltage across the resistor is IR. The voltage across the capacitor
lags the voltage by 90° and the voltage across the inductor leads the current by
90°. The voltage across the resistor is in phase with the current. The pointer
system in Figure 2.10 shows this timing relationship. The voltage across the
resistor is the reference voltage, and it points to the right.
The voltage pointers for the capacitor and inductor point in opposite
directions. This means that these voltages subtract. In Figure 2.10 the voltage
across the inductor dominates and the result is a net inductive reactance. We can
now construct a pointer that is the sum of these three voltages. The difference
voltage IXL - IXc is one side of a rectangle. The other side is IR. The length of
the diagonal is V = VI -'R' + 12(XL - XC)2. The ratio V/I is the impedance of
the circuit or
(Continued)
LEARNING CIRCUIT 10
1 10-mH inductor
1 100-0 resistor
1 0.01-µF capacitor
Turn back and look again at Figure 2.9. The RLC filter in this figure uses a
10-mH inductor, an 0.01-µF capacitor, and a 100-Q resistor. This circuit has a
gain of 1 at dc and low frequencies. As the frequency of the input sine wave
nears the resonant frequency, the current rises as the impedance drops. At
resonance the current is maximum, and the voltage across the capacitor can be
quite large. The voltage is limited by the value of the resistor. Assume the input
voltage is a 1-V sine wave. If the reactance of the capacitor at resonance is 1,000
S2 and the resistor is 100 0, the output voltage will be 10 V. The output voltage
amplitude is maximum near the resonant frequency.
Connect a sine wave generator to these three components and you can
observe this amplitude response across the capacitor. If you were testing a circuit
and you saw this response, it could be a sign of instability, which is not
desirable. In our circuit, if the resistor is increased, the peaking effect is reduced.
In a circuit where the equivalent resistance goes to 0, the result will be an
oscillator.
When the circuit in Figure 2.9 is excited by a square wave at 1 kHz, the result
is ringing. This ringing voltage is shown in Figure 2.12. If the resistor is
increased, the ringing will diminish. There is a point where the square wave
response has 2 or 3 percent of overshoot. This is an optimum response. In
designing circuits where stability is a consideration, this is how the circuit
should respond. This demonstration illustrates the power of a square wave. By
observing the square wave voltage response, we can infer the response of the
circuit to a wide range of sine waves. Change the resistor to 1,000 S2 and
observe that the ringing is reduced. At R = 1,400 S2, the response is optimum.
The inductor in the parallel resonant circuit has some resistance. To see the
effect this resistance has on the circuit impedance, it is convenient to consider
the energy lost in the resistance. As an example, consider 10 V across a parallel
resonant circuit where L = 10 mH and C = 0.10 µF. The resonant frequency is
15.92 kHz. The reactance of the inductor is 1,000 Q. The current that circulates
in the resonant circuit for V = 10 V is 10 mA. If the resistance in the inductor is
10 S2, the power dissipated is hR = 1 mW. This power must be supplied by the
10-V source. The power is VI = 1 mW. But V = 10 V and the current I must
equal 0.1 mA. The impedance given by Ohm's law is V/1= 10/0.0001 = 100 W.
This is a high impedance compared to the reactance. If the series resistor were
only 1 S2, the impedance would increase to 1 M.
LEARNING CIRCUIT 11
Transformers
At frequencies above 1 MHz, the solenoid does not need a core; the magnetic
path can be in air. In applications at power frequencies (60 Hz), however, we
must provide a magnetic material as a path for the magnetic flux.
When a voltage is applied to a coil of wire, the current that flows is called the
magnetizing current. This reactive current is associated with an inductance
called the magnetizing inductance. The presence of iron in the magnetic path
reduces the amount of magnetizing current that is required. If the induction flux
follows a magnetic material along its entire path, the magnetizing current can be
significantly reduced. The reduction factor is known as the permeability of iron.
In the previous example, a permeability of 10,000 would reduce the current from
160 A to 16 mA, an acceptable level.
The induction flux created by a voltage across a coil of wire follows the
magnetic path as this path stores the least amount of magnetic field energy. If a
second coil is wrapped around the first coil, then any changing induction flux
threads both coils. The voltage on this added coil depends on how rapidly the
induction flux is changing and on the number of turns. If both coils have 1,000
turns, then the voltage on both coils will be identical. If the second coil has 500
turns, the voltage will be half. This transfer of voltages between coils of wire is
called transformer action. The coil receiving the initial voltage is called the
primary coil. All other coils are called secondary coils.
The symbol for a transformer is shown in Figure 2.14. The bars between the
coils represent the iron in the core. This symbol is sometimes misleading,
because it does not represent the actual construction of the transformer. For
example, there is capacitance between the primary coil and the secondary coil.
The capacitance is not symmetrically distributed, as the symbol might suggest,
but is largely between the last turns of the primary coil and the first turns of the
secondary coil.
Power transformers (60 Hz) are built by winding the coils on a bobbin. The
core material is then added by interleaving iron laminations. These laminations
provide a magnetic path for the magnetic flux. Lam inations are required
because the magnetic field at 60 Hz can only penetrate a short distance into iron.
A typical lamination thickness at 60 Hz is 15 mils or 0.015 inches. 1 mil is a
thousandth of an inch. At 400 Hz the lamination thickness is about 5 mils.
When the number of field lines that thread through a one turn of wire
increases at a linear rate, a steady voltage will appear at the ends of the turn.
When a coil of wire replaces the single loop, the voltage is proportional to the
number of turns in the coil. When a voltage is placed on a single turn of wire, the
number of field lines that thread the coil must change at a fixed rate. If the
applied voltage is fixed and if the number of turns is increased, the voltage per
turn is decreased. This reduces the rate at which the field lines must change.
If the number of turns in the coil were decreased to 250, the number of field
lines would have to change at twice the rate. This means the maximum number
of lines per cm2 would be 19,200. This is enough to saturate the magnetic
material. To avoid saturation the core area could be increased or the number of
turns increased.
The secondary voltages of a transformer are often lower than the primary
voltage. The ratio of secondary turns to primary turns determines the voltage. If
the primary has 1,000 turns and is 120 V, a secondary coil of 100 turns will be
12 V. If a load is placed on this secondary coil, current will flow as determined
by Ohm's law. For example, if the load is 12 S2, the current is 1 A. The power
dissipated is 12 W. This means that the 120-V primary voltage must supply 0.1
A to the primary coil. Viewed from the primary side of the transformer, a
voltage of 120 V and 0.1 A represents an impedance of 1,200 Q. A 12-Q load on
the secondary represents a 1,200-52 load to the primary voltage. The ratio of
load impedance to input impedance is 100:1. This is the square of the turns ratio.
If a transformer has two secondary coils, then each coil can be independently
loaded. The current waveforms demanded by the secondary coils are reflected on
the primary by the turns ratio. For example, if the primary voltage is 120 V and
the secondary voltage is 12 V, a 3-A pulse of current on the secondary is
supplied by a 0.3-A pulse of current in the primary coil.
The resistances of the primary and secondary coils are adjusted by the
designer so that power losses in the coils are about equally divided. Since the
primary coil must handle all the power, this coil occupies about half the
available volume. It is common practice to wind the primary coil next to the
core. Coils that are intended to supply a higher current have the lowest
resistance.
Many small power transformers have high coil resistances. The result is that
magnetizing current flowing in the resistance of the primary coil modifies the
voltage waveform appearing on the primary coil. This change in voltage
waveform appears on the secondary coils. The magnetizing current is apt to be
greatest when the voltage waveform is at a zero crossing. In most dc power
supplies the current is supplied in short pulses. This current flows in the coil
resistances of the transformer and further reduces the voltage available to the
load. In short, small power transformers (up to 10 W) have many practical
limitations. The loaded waveforms are often very different than sine waves.
Diodes
A diode is a component that allows current to flow in one direction (see chapter
8). The symbol for a diode is
The current flows in the direction of the arrow This is called the forward
direction of the diode. In the reverse direction a diode will withstand a high
voltage before conducting. There are many types of diodes available to
accommodate high current, high reverse voltage, and high-frequency
applications. The forward drop in a silicon diode is about 0.6 V. Diodes are often
used in groups of four. This configuration is known as a bridge rectifier A
package of four diodes is available as a single component with four connections.
We will discuss the diode in detail in the next chapter. For the moment, we need
the diode to discuss power supplies.
DC Power Supplies
Most of the circuits in electronics function from dc power supplies, but batteries
are the sole voltage source in some applications. Sometimes rechargeable
batteries are used, and the batteries are recharged whenever utility power is
available. In this section we will look at dc voltages derived from utility power,
not batteries.
The output voltage V follows the ac source voltage on the positive half of
the power cycle. A 10,000-52 resistor is used as a load so that the voltage is 0
when the diode is not conducting. This is called half-wave rectification. If the
adapter is rated 10 V unloaded, the peak output voltage is 14.14 V less a diode
drop of 0.6 V. This is approximately 13.54 V.
This half sine wave can be used to charge a capacitor. This added capacitor
is shown in Figure 2.16.
The capacitor charges to the peak ac voltage, but the capacitor cannot
discharge back into the transformer as the diode prevents current flow in that
direction. Load resistor RL draws current from the capacitor, causing the
voltage to sag each cycle. The ac voltage rises in each cycle to recharge the
capacitor. The charging current flows when transformer voltage exceeds the
capacitor voltage plus the diode voltage drop.
Figure 2.15 The action of a diode on a sine wave voltage
Figure 2.16 A diode and a capacitor providing a dc voltage
The problem with the circuit in Figure 2.16 is that a dc current flows in the
transformer secondary winding. This will tend to saturate the transformer
core, and the transformer may draw excessive magnetizing current. This
excessive current can overheat the transformer. To avoid this problem, a
second diode can be used to supply a second power supply. If the transformer
supplies equal current on the other halfcycle, the transformer will not saturate.
A power supply with a positive and negative dc voltage finds wide acceptance
in circuit design. This circuit is shown in Figure 2.17.
Figure 2.17 The use of two diodes to form two dc power supply voltages
LEARNING CIRCUIT 12
Building a DC Power Supply
circuit board
bus wire
2 1N1002 diodes
(Continued)
The two dc voltages are the ac voltage times 1.414, less a diode drop of 0.6 V.
This means the dc voltages should be about 25 V plus and minus.
Observe the dc voltages using the oscilloscope with the input shield
connected to the 0 of potential or common. This is the conductor between the
two capacitors. Place 1,000-52 1-W resistors between the two dc voltages and
common and observe that the dc voltage drops slightly. Place a second 1,000-52
1-W resistor across each supply and again note the voltage change. Calculate the
source impedance. Calculate the power dissipated in each resistor. Notice the
ripple voltage (peak-to-peak ac voltage) across the capacitors for one set of
resistors and for the second set. To observe power supply ripple voltage on your
oscilloscope, it may have to be in the ac coupling mode.
When you build this power supply, do so on one end of the circuit board and
leave room for other circuits. We will be using it in future Learning Circuit
exercises. For ease in adding later circuits, use three long conductors that go
across the board. The top conductor is +25 V, the middle conductor is 0 V
(common or ground), and the bottom conductor is -25 V.
You may have observed that Figures 2.17 and 2.18 are very different drawings,
yet they both describe the same circuit. Figure 2.17 is what is called a schematic.
It shows only the actual components used in the circuit. Figure 2.18 is a
construction diagram, and shows how these circuits could be actually
constructed by being laid out on a circuit board. It is only one of many possible
layouts or arrangements. The schematic is unique to the circuit; the construction
layout is simply a suggestion of a possible way the circuit might be constructed.
In several previous drawings (e.g., Figures 2.9 and 2.13), I drew them both on
the same page, with the schematic above and the construction diagram below.
Now that you are building larger and more complicated circuits, I will need to
provide you with separate drawings-there will not be enough space to include
them both in one figure.
The circuit in Figure 2.19 can be used without a transformer centertap. The
two capacitors are then replaced by one capacitor that is rated for the full
voltage. The four diodes are the full-wave bridge rectifier mentioned earlier. The
transformer is connected to bridge terminals marked ac, and the other two
terminals are marked plus and minus. When a full-bridge rectifier is used in this
configuration, the current must flow through two diodes to charge the capacitor
on each halfcycle.
The circuit in Figure 2.16 shows that a series diode and capacitor rectify the
voltage and a dc voltage appears across the capacitor. If the positions of the
capacitor and the diode are reversed, a dc voltage will still appear across the
capacitor. One side of the capacitor is connected to the ac voltage source. The
other side of the capacitor has this same ac voltage plus a dc value. If a second
capacitor and diode are added, the result is rectification of the ac voltage
superposed on the dc output of the first diode and capacitor. The resulting
voltage is double the first dc voltage. This circuit is shown in Figure 2.20.
LEARNING CIRCUIT 13
3 1.0-gF capacitors
2 1N1002 diodes
1 100-kU resistor
1 100-kU resistor
SELF-TEST
2. 10 V dc is impressed across a 2-mH inductor. How long does it take for the
current to reach 100 mA?
6. Inductors of 1 mH, 2 mH, 3 mH, and 4 mH can be switched to provide all the
values from 1 mH to 10 mH. Is there another group of four inductors you can
use?
9. A relay coil has an inductance of 2 H and a coil resistance of 200 Q. How long
does it take for the current to rise to 63% of final value?
11. If the source voltage in problem 8 is 10 V, what is the current in one time
constant?
12. In problem 10, what is the phase angle between the source voltage and the
current?
14. A 100-52 resistor, a 10-mH inductor, and a 0.01-µF capacitor are in series.
What is the resonant frequency?
15. In problem 14, what is the reactance of the capacitor at this frequency?
16. In problem 15, what is the reactance of the inductor at one-half the natural
frequency?
17. In problem 15, what is the reactance of the capacitor at one-half the resonant
frequency?
20. What is the impedance of the circuit in problem 14 at twice the resonant
frequency?
21. A 1-mH inductor and a 500-pF capacitor form a parallel resonant circuit.
What is the resonant frequency?
23. In problem 21, the resistance of the inductor is 20 Q. What is the impedance
of the circuit?
24. The open circuit voltages on two coils of a 120-V transformer are 25 V and 8
V. What are the turns ratios?
25. The load resistors in problem 24 are 50 S2 and 16 Q. What is the primary
current? Neglect magnetizing current.
29. A transformer secondary is marked 15-0-15 V. What are the two dc voltages
that can be supplied?
30. A full-wave bridge is placed across the full voltage in problem 29. What is
the output dc filtered voltage?
31. A 10-V dc voltage sags 0.5 V in each halfcycle of 60 Hz. The current
averages 100 mA. What is the size of the filter capacitor?
ANSWERS
4. XX=62852.
5. X.=31.4 Q.
6. 1,2,4,8.
7. % µH.
9. L/R = 0.01 s.
12. 45°.
13. 10 W.
15. 10,000 Q.
16. 5,000 Q.
17. 20,000 Q.
18. 1000.
19. 15 W.
20. 15 W.
24. If the primary has 1,200 turns, the secondaries are 250 and 150 turns.
26. !2R=0.075 W.
• how diodes and zener diodes are used to clamp and limit signals
Silicon dioxide (the element silicon combined with oxygen) is the mineral
quartz, which makes up most of the rock and sand on Earth. Pure silicon is rare
in nature but can be grown as a crystal. A pure silicon crystal is an insulator. But
when a very small percentage of impurity atoms are added to the crystal, it
becomes a conductor. This is why it is called a semiconductor.
When boron is added as a dopant, the silicon is called p-type, as the boron
atom provides spaces (holes) that electrons can use to move through the crystal.
P-type silicon is also a conductor.
Diodes
This limiting of current to one direction is the basis of diode action, which is
also called rectification. Figure 3.1 shows the voltage/current curve for a typical
silicon diode. Notice that a forward voltage of about 0.6 V is required before
current flows.
One application of a diode is a voltage clamp. The diode acts like a switch
contact when it is conducting in the forward direction. The diode can act to
clamp a signal so that it cannot change. When a voltage is clamped, the diode
voltage drop is approximately 0.6 V whether the current is 1 mA or 1 A. In the
reverse direction there is no current flow for a voltage difference of 1, 10, or 100
V. This is the characteristic of an open circuit. Figures 3.2 and 3.3 show a circuit
that allows voltages greater than -0.6 V to pass and voltages less than -0.6 V to
be blocked (clamped).
LEARNING CIRCUIT 14
You will need (in addition to your circuit board and measuring
equipment) :
1 1-kQ resistor
1 1N4148 diode
1. Connect the diode and the resistor as shown in Figures 3.2 and 3.3 to
the sine wave output of a function generator.
3. Reverse the diode and note that the positive voltage is clamped.
Note that from this point forward in the book, when two consecutive figures
show a Learning Circuit, the schematic drawing will be given first and the
construction diagram will follow.
In this circuit, the diode conducts only when the signal is more negative than
0.6 V relative to the common or reference conductor. If the diode is reversed in
direction, the circuit allows only voltages less than +0.6 V to pass.
When a diode clamps a voltage source, it acts as a short circuit. This short
circuit can cause overload or overheating if the source is a low impedance. In
this case a resistor is usually placed between the voltage source and the diode. If
the maximum voltage that can be tolerated is 10 V and the maximum current is
10 mA, the resistor must be 1,000 SZ or greater.
Diodes are often used to limit a signal voltage so it cannot exceed the power
supply voltages. This is done to protect the circuit against damage. In this case
the diode clamps are connected between the signal and the power supply
voltages. This circuit is shown in Figures 3.4 and 3.5. If the signal voltage is
between the limits of the power supply voltages, the signal is unaffected
(unclamped).
LEARNING CIRCUIT 15
You will need (in addition to your circuit board and measuring
equipment) :
2 9-V batteries
2 1N4148 diodes
1 1-ku resistor
Connect the circuit in Figures 3.4 and 3.5. Use a sine wave generator
to show that voltages above 9.6 V are clamped.
Figure 3.4 Diode clamps used to block signal voltages greater than the power
supply voltages
Figure 3.5 The construction of the circuit in Figure 3.4
Zener diode voltages do change with temperature. The most stable range is
around 5.6 V. Zener diodes in this voltage range are often used as voltage
references because of their stability.
Zeller diodes should never be placed directly across a power supply. If the
voltage exceeds the breakdown voltage, there is a chance that the zener diode
will be destroyed or the power supply will be damaged.
Source Impedance
The voltage across a zener diode is nearly constant over a wide range of current.
This makes a zener diode useful as a voltage regulator. Figures 3.7 and 3.8 show
an application where a source voltage that varies from 9 to 18 V provides current
for a 5.1-V zener diode. The voltage across the zener diode varies less than 0.05
V. The zener voltage changes only 1% for a 100% change in source voltage.
This is a regulation factor of 100.
Figure 3.7 A zener diode voltage regulator
Figure 3.8 The construction of the circuit in Figure 3.7
When a zener diode is used as a power supply, any added load current is
taken from the zener diode current. Assume that in Figure 3.7 the unloaded zener
diode current is 10 mA. If the added load resistor is 1,000 S2, the zener current
drops to 5 mA.
A zener diode can be used as a clamp so that a signal cannot exceed the zener
voltage. In the negative direction the forward direction of the zener clamps the
signal as a standard diode. To limit the voltage symmetrically, two zener diodes
can be used back to back. This circuit is shown in Figure 3.9. If the zener
voltages are 5.1 V, the clamping voltage is actually 5.7 V (the zener voltage plus
a diode drop of 0.6 V). In this example the signal voltage cannot exceed ±5.7 V.
Figure 3.9 Two zener diodes acting as a signal clamp
LEARNING CIRCUIT 16
Observing Voltage Regulation Using a Zener Diode
You will need (in addition to your circuit board and measuring equipment):
2 9-V batteries
1 1-ku resistor
1. Connect the circuit shown in Figures 3.7 and 3.8, using two 9-V batteries in
series to supply current to a 5.1-V zener diode.
3. Now use one of the batteries and note the zener voltage. What was the
regulation factor? The bar on the diode corresponds to the bar in the
schematic.
LEARNING CIRCUIT 17
You will need (in addition to your circuit board and measuring equipment) :
1 1-ku resistor
Construct the circuit shown in Figures 3.9 and 3.10, placing two 5.1-V zener
diodes back to back along with a resistor across a sine wave voltage source.
Observe that the signal voltage is limited by the zener voltage to ±5.7 V.
Figure 3.10 The construction of the circuit in Figure
3.9
We will now take up the subject of transistors. Transistor are called active
elements, as opposed to resistors, capacitors, and inductors, which are passive
elements. Passive elements simply react to a signal without enhancing it, but
active elements like transistors use input from a power supply to enhance a
signal in some way. Diodes and zener diodes are considered nonlinear passive
elements, which means that they change the waveform but do not otherwise add
to it.
The outer n layers are called the emitter and the collector. The p material is
called the base. The two n layers may appear to be symmetrical, but they are not;
they must be connected in the correct order as specified by the manufacturer. If
you interchange the emitter and the collector, you may damage the transistor. If
the base is disconnected, the pnp layers form two back-to-back diodes, which
means that current is theoretically blocked and cannot flow in either direction.
The relationship between base current, collector current, and collector voltage
in a typical npn transistor is shown in Figure 3.12. The individual curves
represent a fixed base current. Notice that the collector current is controlled by
the base current and not by the collector voltage.
Transistor Gain
Transistors that operate along one of the curves in Figure 3.12 are operating in
their linear range. This is the operating range we will consider in this section.
There is a distinction between operating voltages and currents and signal
voltages and currents. A signal makes changes to the base current and voltage.
These changes result in new collector voltages and currents. The ratio of changes
to two parameters is called gain.
In most circuits it is desirable to have a voltage gain, not a current gain. The
transistor must be a part of a circuit so that the current gain ((3) of a transistor
can provide this voltage gain. Changes in collector current can be converted to
changes in voltage by using a collector resistor. Controlling the voltage
difference between the base and emitter is not as easy. To be effective, a circuit
must automatically set this difference voltage so that the transistor is operating in
its linear range. Attempts to obtain gain from a transistor by applying a voltage
between the base and the emitter will usually result in failure.
The circuit in Figure 3.13 is called an emitter follower. The voltage between the
base and emitter is automatically set to provide the required collector current.
When the base moves one volt, the emitter follows by almost one volt. This
means the gain of the emitter follower circuit is very close to unity. This is a
good time to define the term voltage gain. Voltage gain is the ratio of output
voltage change to input voltage change.
In the circuit of Figure 3.13, assume the input base voltage is set to +2 V. The
emitter voltage follows and is at +1.4 V. The new emitter current is 11.4/2,700 =
4.2 mA. The new base current is 21.1 .iA. This means that the base has moved 1
V and the base current has changed 3.7.tA. The ratio between voltage change
and current change is called input impedance.
LEARNING CIRCUIT 18
You will need (in addition to your circuit board and test
equipment):
1 2.7-ku resistor
In the circuit shown in Figure 3.13, the ratio of changes is V/I = 1/3.7 x 10-6
Q = 270,000 0. (This measure assumes the 10-ki2 resistor is not present.) If the
transistor 0 is 400, the input impedance would be 540,000 Q. The input
resistance of a transistor that is not used as an emitter follower is typically
around 100 Q. (Note that this is an example of when you are likely to hear the
term impedance used in place of resistance. Impedance is incorrect, since sine
waves were not used in the measurement. However, in this circuit, measurement
using sine wave voltages would give the same result.)
In Figure 3.15, assume the base voltage has been adjusted so that the emitter
voltage is +5 V. A 100-52 load resistor R is placed between this emitter and the
common lead. The current in the resistor must flow in the transistor. This means
the base current must increase slightly. The emitter voltage might drop 0.05 V to
4.95 V.
In this example the voltage change is 0.05 V and the current change is 50 mA.
The output impedance is V/I = 1 Q. This example shows that an emitter follower
has a low output impedance.
Emitter followers have a gain very close to 1, a high input impedance, and a
low output impedance. In effect, an emitter follower is an impedance converter.
A voltage source with a high source impedance or with a limited ability to
supply current can be converted to a voltage source with a low source impedance
that can supply current. In our example, the input current is in the microampere
range and the output current is in the milliampere range. This multiplication of
current is one form of gain.
Figure 3.15 An emitter follower with a load resistor
Figure 3.16 The construction of the circuit in Figure 3.15
The emitter follower follows an input signal no matter how slowly it changes.
Amplifiers that handle this type of signal are called dc amplifiers. An emitter
follower is a dc amplifier with a gain of 1. If the input voltage changes 1 V at dc,
the output voltage changes 1 V at dc. The 0.6-V voltage difference is an offset
signal. In applications such as audio amplifiers where only ac is of interest, a
capacitor can block the offset voltage. This capacitor and any terminating
resistor form a high-pass filter. The -3 dB point occurs at the frequency where
the reactance of the capacitor equals the terminating resistor.
LEARNING CIRCUIT 19
You will need (in addition to your circuit board and test equipment):
1. Use your power supply and the emitter follower you built in Learning Circuit
18.
3. Measure the emitter voltage with respect to the common lead. It should read
about 4.4 V.
4. Place a 100-1 resistor from the emitter to common. This is resistor R in Figure
3.16.
5. Measure the emitter voltage. What is the output impedance? Save this circuit
construction.
LEARNING CIRCUIT 20
LEARNING CIRCUIT 20
1. Use the circuit you built in Learning Circuit 19. Measure the input base
voltage and the emitter output voltage with respect to power supply common.
2. Use a clip lead to connect the base to the power supply common (this removes
the 5-V input base signal) and measure the emitter voltage. Use these voltage
measurements to determine the gain.
3. Remove the 100-0 resistor load from the emitter to the common.
4. Repeat these steps to obtain new voltages. What is the new gain?
The emitter follower in Figure 3.15 can supply current to a load resistor.
When the input voltage is positive, the output voltage follows. The load current
is added to the collector current. In the negative voltage direction, the current in
the load is subtracted from the collector current. In the negative direction the
emitter follower cannot support a load current that exceeds the nominal collector
current. The collector current for various output voltages is shown in Figure
3.17. In the next chapter we will discuss a circuit that does not have this
limitation.
The emitter follower circuit using a single voltage power supply is shown in
Figure 3.18. A voltage divider must be used to establish a base voltage so that
the transistor can draw current. This transistor current is determined by the base
voltage and the value of the emitter resistor. The voltage divider must supply any
required base current.
Figure 3.17 The current and voltage in an emitter follower
It is a good idea for the voltage divider current in Figure 3.18 to be 10 times
the base current. If the base current is 100 µA, the divider current can be 1 mA.
Two coupling capacitors are required if the operating voltages are to be blocked.
In this circuit the input impedance is dominated by the voltage divider. This
impedance is equal to the two divider resistors in parallel. The value of the
output capacitor depends on the terminating resistor, not on the source
impedance.
You will see in chapter 5 that an integrated circuit can be used to build the
equivalent of an emitter follower. This approach is often less expensive and has
the additional advantage of having no offset problems. Emitter followers are still
used when the output current must be greater than that supplied by an IC
amplifier, for example in supplying current to a loudspeaker.
Figure 3.18 An emitter follower and a single voltage power supply
The circuit in Figure 3.19 provides a negative gain of 3.0. In this circuit we make
use of a dual 25-V power supply, two zener diodes, resistor R, = 3.3 M, and
resistor R2 = 10,000 Q.
The two zener diodes are used to provide a -5.1-V power supply voltage. The
transistor current can vary, but the zener voltage stays con stant. When the base
voltage is 0, the emitter voltage is approximately -0.6 V. The voltage across the
emitter resistor is 4.4 V. The current in the emitter resistor is 4.4 V divided by R,
= 1.333 mA. This current flows in the collector resistor R2. The voltage drop in
this resistor is IR2 = 13.33 V. If the power supply voltage is 25 V, the collector
voltage is 11.7 V. These are operating voltages.
Figure 3.19 A voltage gain using an npn transistor
If the input voltage rises 1.0 V, the new emitter voltage is +0.4 V. The current
in the emitter resistor increases to 1.8 mA. This current flows in the collector
resistor. The resulting voltage drop is 18 V. The new collector voltage is 8.7 V.
The collector voltage dropped 3 V. The gain is the ratio of output voltage change
to input voltage change, or -3 V/ 1.0 V = -3.00. This is also the ratio of R2/R,. If
the base voltage changes sinusoidally, the output voltage also changes
sinusoidally. When the sine wave has a peak value of +1.0 V, the output signal
has a peak voltage of -3 V.
LEARNING CIRCUIT 21
Obtaining Voltage Gain from a Transistor
You will need (in addition to your circuit board and test
equipment):
1 2N3904 transistor
1. Build the circuit shown in Figures 3.19 and 3.20, using the
circuit values indicated. Be careful to connect the transistor
correctly. A diagram of the transistor pin layout is shown in
the upper right corner of Figure 3.20. Check that the dc
voltages are correct.
7. Turn on the power and verify that the circuit gain is now
approximately 5.
Voltage Gain at AC
The circuit in Figure 3.19 can be modified to have a gain of 10 at ac but a gain of
less than 1 at dc. This circuit is shown in Figures 3.21 and 3.22. The collector
resistor is set to 10 W. The emitter resistance is made up of two resistors: R, = 1
kS and R3 = 22 W.
In Figure 3.19 the zener diode held the emitter resistor return to -5 V at all
frequencies. In Figure 3.21 a capacitor holds the emitter resistor return fixed at
frequencies above 150 Hz. The advantage of this circuit is that the dc operating
conditions do not vary, as there is no gain at dc. This technique is used in high-
frequency amplifiers where dc gain is unimportant and stable operating
conditions are desirable.
LEARNING CIRCUIT 22
You will need (in addition to your circuit board and test equipment):
1 2N3904 transistor
1 1-µF capacitor
Figure 3.23 shows the structure and symbol for a pnp transistor. The outer two
layers of p-type silicon are called the emitter and the collector and the center n
silicon layer is called the base. This pnp transistor is very similar to the npn
transistor except that all the voltages are reversed. In the next section the pnp
transistor is used as an emitter follower.
Note that the offset voltage is +0.6 V instead of -0.6 V. In this circuit any
output load current for a negative output voltage is added to the transistor
current. In the positive direction the output load current subtracts from the
collector current. When the transistor current is zero the output voltage is at its
maximum.
LEARNING CIRCUIT 23
LEARNING CIRCUIT 23
1 2N3906 transistor
Using Learning Circuit 22, modify the circuit to that shown in Figures 3.24
and 3.25. Go through the same steps as before. The pnp and npn transistors have
the same pin arrangement. Verify that the operating voltages are correct. Before
connecting the signal generator, verify that the signal level is 1 V peak-to peak.
Measure the ac gain with and without a load resistor of 100 Q.
Figure 3.25 The construction of the circuit in Figure
3.24
The convention of placing the positive power supply conductor at the top of
the page means that current always appears to flow from top to bottom. The
arrows on the emitters of npn and pnp transistors point in the direction of current
flow and thus they always point toward the bottom of the page. These
conventions make it easier to read a schematic.
The base current for the pnp transistor is supplied by the collector of the
npn transistor. If the (3 of the pnp transistor is 200, this base cur rent is a small
fraction of the collector current. The overall gain of this circuit is +9.2. If the
first stage emitter resistor is shunted by 10 kS2, the overall gain will be very
nearly 10.
Figure 3.26 An npn and pnp transistor used to provide gain
Figure 3.27 The construction of the circuit in Figure 3.26
The circuit of Figure 3.26 may have an offset of several volts. This offset can
be removed by injecting a small dc signal into the base of the second stage. A
zeroing control is provided, as the amount of offset is unknown. The control
component is called a potentiometer. It is a resistor with a slider that makes
contact along the resistance. The potentiometer forms an adjustable voltage
attenuator. In this circuit the slider provides any voltage from +25 V to -25 V.
This variable voltage changes the base current to the second stage and corrects
for the offset. Later, when we use an IC amplifier, this adjustment may not be
necessary. Offsets and dc drift are always a problem with circuits that amplify
both slowly and rapidly changing voltages. This is another example of a dc
amplifier.
LEARNING CIRCUIT 24
You will need (in addition to your circuit board and test equipment):
1 10-ku potentiometer
3. Measure the gain of the circuit using the oscilloscope. If the offset is
a problem, adjust the potentiometer to set the output voltage to 0.
SELF-TEST
1. The current in a 15-V zener diode is 10 mA. What is the power dissipation?
2. If the load resistor across the zener diode in problem 1 is 3 kS2, what is the
power dissipation in the zener diode?
3. The voltage across a zener diode is 9.75 V. When a 5-ku resistor is placed
across the diode, the voltage drops to 9.60 V. What is the source impedance?
5. The input to an emitter follower changes from -0.5 to 1.5 V. The emitter
changes from -1.2 to 0.75 V. What is the gain of the emitter follower?
6. The input current in problem 5 changes from 0.8 mA to 0.9 mA. What is the
input impedance?
10. The power voltage varies from 105 V to 125 V. An emitter follower requires
a minimum of 5 V dc from the collector to the emitter. A dual power supply is
used to supply voltages for an emitter follower. The emitter follower is used
to regulate 15 V dc. Allow for one diode drop and a 2-V drop in the
transformer coil plus a peak-to-peak ripple voltage of 1 V. What should the
unloaded rms voltage be for the secondary of the transformer for a nominal
voltage of 117 V? Hint: Add up the voltage drops at the lowest power line
voltage. Determine the minimum secondary voltage. Then correct for the line
voltage.
ANSWERS
1. 150 mW.
2. The current in the load is 5 mA. The zener current is reduced to 5 mA. The
dissipation is 75 mW.
3. The change in voltage is 0.15 V. The change in current is 1.92 mA. The
source impedance is 16.3 Q.
4. The collector current changes 0.5 mA. The base current changes 2 NA. The (i
is 250.
5. The output changes 1.95 V when the input changes 2 V. The gain is 0.975.
7. 3.4 Q.
8. Both loads take 118 mA. The total current is 138 mA.
10. At the lowest line voltage the secondary voltage must be 15 V + 5 V + 0.6 V
+ 2 V + 1 V = 23.6 V. The line voltage correction is 117/105 = 1.114. The
secondary voltage should be set at 26.3 V.
Objectives
Most amplifiers or circuits with gain provide output voltage to various loads.
These loads could be a speaker or a long cable. As we saw earlier, the simple
npn emitter follower must draw a steady average current in order to supply a
negative output voltage to a load. For a negative output voltage, the load current
is subtracted from the collector current. This is a drawback, but it can be avoided
by using two emitter followers, as shown in Figures 4.1 and 4.2.
The top transistor is an npn type and the bottom transistor is a pnp type. The
top transistor supplies load current when the output voltage is positive, and the
bottom transistor supplies load current when the output voltage is negative. The
static current through the two transistors when the output voltage is 0 can be
limited to a few milliamperes. The symmetry of this circuit allows the output
voltage to be 0 when the input voltage is 0.
The operating voltage at the bases of the two transistors can be set so that a
small amount of collector current flows when the output is at 0 volts. This is
accomplished by a voltage divider using two low-current signal diodes and
resistors R, and R2. The forward drop in these diodes is sufficient to set the base
voltages to +0.6 V and -0.6 V. The midpoint of the diodes is the input terminal.
This input point is normally connected to a signal from another circuit. In the
Learning Circuit I have placed a resistor of 10 kS2 between the input and the
circuit common.
Figure 4.1 A stacked emitter follower
Figure 4.2 The construction of the circuit in Figure 4.1
The two 10-Q emitter resistors R3 and R4 are required to avoid a possible
instability. It is good practice in electronics to avoid paralleling connections
between two or more active elements. In this circuit the transistor bases are
connected together though diodes. Tying the two emitters together would violate
this rule. The 10-Q resistors in the emitters are a safety factor to remove any
possible instability. The instability could be an oscillation above 20 MHz, which
could overheat the transistors and yet go undetected.
The input impedance of the emitter followers is high compared to the
impedance of the voltage divider. If the I of the output transistors is 150 and the
maximum transistor current is 100 mA, the maximum base current is 0.66 mA.
The voltage divider current should be about 6 mA. If the divider resistors are 3.3
kS2, the input impedance is approximately 1,600 Q. As we will see later, a
standard IC amplifier can easily supply current to this load. A signal generator
with a peak-to-peak output of 20 V can drive this circuit to show its output
capability.
An ideal constant voltage source is a voltage that does not vary with a change in
load current or a change in power supply voltages. In prac tice, voltage
regulators are not perfect. In many situations regulation to within 5% is totally
acceptable.
LEARNING CIRCUIT 25
LEARNING CIRCUIT 26
1 TIP29A transistor
3. Measure the power supply ripple at the collector of the pass element
(25 V). Note the power supply ripple at the output of the regulator.
Most of it should be removed. The voltage should not shift more than
a tenth of a volt; if it does, check over your connections carefully to
be sure they are correct.
The second example of a voltage source was the emitter follower in Figure
3.15. Here the emitter output impedance was low If the base input voltage is
fixed, the emitter output voltage is also nearly fixed. If you give this same circuit
a fixed input voltage, it becomes a positive voltage regulator, as shown in Figure
4.3.
Figure 4.3 A positive voltage regulator
LEARNING CIRCUIT 27
You will need (in addition to a power supply and measuring equipment) :
1 TIP30A transistor
Add the circuit of Figure 4.5 to your power supply. The construction is shown
in the bottom half of Figure 4.4. Now you have a regulated plus and minus 15-V
power supply. This negative voltage regulator should be tested the same way the
positive regulator was tested in Learning Circuit 26. Note the voltage at the
emitter follower before and after applying a 330-52 load resistor. Compare the
ripple reduction from the collector to the emitter on the pass element.
Remember, ripple voltage is the ac component riding in the dc power supply
voltage.
The zener current is supplied from the unregulated voltage source through a
resistor. This zener voltage defines the base voltage of the transistor. The emitter
of the transistor is the output voltage. The transistor functions as an emitter
follower, except with a fixed base voltage. When a TIP30A transistor is used,
this type of voltage regulator can easily supply 100 mA. The transistor is often
called a pass element, as the load current must pass through the transistor.
The third example of a voltage source was the pnp emitter follower in Figure
3.24. If the base voltage is defined by a zener diode, the output of the emitter
follower is regulated. Again the zener current is supplied from the unregulated
voltage source through a resistor. The transistor functions as an emitter follower
except that the base voltage is fixed. This circuit is shown in Figure 4.5. With a
TIP30A transistor, this voltage regulator can easily supply 100 mA. This
transistor is also called a pass element.
Figure 4.4 The construction of the circuits in Figures 4.3 and 4.5
Figure 4.5 A negative voltage regulator
Voltage Sources
The ideal voltage source supplies a voltage that is independent of the load
current. The voltage source might be a dc value, or any waveform for that
matter. The load could be a resistor, an open circuit, or even a capacitor. If the
voltage waveform is a sine wave, then the concept of output impedance can be
used. In all practical circuits the regulation varies with frequency. If the source
impedance measures 1 S2 from dc to 100 kHz and is 10 S2, at 1 MHz, then the
source impedance is inductive. There is no physical inductor. The circuit simply
performs as if an inductor were present. In this example the inductance has a
reactance of about 10 S2 at 1 MHz.
Current Sources
Current sources are not as common as voltage sources. A current source supplies
a current that is independent of the load impedance. The load in this case might
be a short circuit or a resistor. A constant current source cannot function into an
open circuit, as the current cannot flow. The higher the resistance value, the
higher the voltage must be for the same current. If the circuit cannot supply the
voltage, the circuit cannot function. As an example, assume a constant current of
10 mA. If 10 mA flows in 100 S2, the voltage is 1 V. If the resistance is 1,000
S2, the voltage is 10 V. If the resistance is 10,000 Q, most circuits cannot
function.
Two constant current circuits that are often used in design are shown in Figures
4.6 and 4.7. The first circuit uses an npn transistor to supply a constant current
from the positive supply to the common. The second circuit supplies a constant
current from the negative supply to the common. In both circuits an emitter
resistor and the zener diode voltage determine the current level. If the resistor is
470 S2, and the zener voltage is 5.1 V, the current is about 10 mA. This current
does not depend on the collector voltage, as long as the transistor is within its
normal operating range.
A constant current source is not affected by any voltages that are in series
with the load. If the resistor is terminated on an ac voltage source at 60 Hz, the
current in the resistor is still constant. The 60-Hz voltage cannot add current to a
current source. This means that the voltage across the resistor will not have any
60-Hz content.
Voltage sources have a similar quality. A voltage source holds constant for
any 60-Hz current that flows in the load. An external source of 60-Hz current
cannot change the voltage from a regulated voltage source. Both of these
statements are true as long as the circuits are within their normal range of
operation.
Figure 4.6 Two constant current circuits
The constant current circuits shown in Figure 4.6 cannot be placed in series.
Any slight imbalance in the current levels will upset the circuits. To get around
this problem, a load resistor must provide a path for the difference current. There
is a similar problem with voltage sources. Voltage sources cannot be placed in
parallel without a connecting resistor. This resistor will limit current flow
between the voltage sources based on the difference voltage. Without this
resistor, voltage sources will overload or self-destruct. To review, current
sources cannot be placed in series and voltage sources cannot be placed in
parallel.
(Continued)
LEARNING CIRCUIT 28
You will need (in addition to a power supply and measuring equipment):
On your power supply board you now have two regulated voltages and a
stacked emitter follower. Now you need room for a constant current source. You
are going to use the two regulated voltages as the power supply for the constant
current sources. Use the circuits in Figure 4.6 and 4.7. The current for the zener
diodes comes from the 25-V power supplies. The collector resistors are 1,000
92. Measure the voltage drop across these resistors. Each current source has been
set up to be 10 mA. The voltage drop across the 1,000 S2 is 10 V. Now change
the resistor to 470 S2 and note that the voltage changes to 4.7 V. What happens
if the resistor is 2,000 fl? Can the resistor be 3,000 SZ?
We will call the transistor base on the left the input or first base. The two
emitters are connected to a constant current source, Q3. In this example the two
base voltages are connected to the zero reference potential through 10-ku
resistors. If the transistors are well matched, then half the available constant
current flows in each transistor. The voltage across each collector resistor is
equal to half the constant current times the resistance value. The collector
voltage is equal to the supply voltage minus this voltage drop. In Figure 4.8 the
total current is 1.0 mA. The collector resistors are 10 kS2s. The voltage drop
across each resistor is 5 V. If the supply voltage is 15 V, the collector voltages
are each 10 V.
The input base current is equal to the collector current divided by the beta of
the transistor. If the beta is 250, then the base current is 0.5mA/250 = 2 .tA. This
current must be supplied or the transistors cannot function. In this example the
input base could be supplied base current through a resistor connected to the
power supply. This resistor is R = 15/2 x 10-6 = 7.5 MQ. In most applications
this base current is supplied by the circuits that connect to the bases.
If the first or input base voltage is raised to 0.01 V and the second base is held
at 0 V, the emitter voltage splits the different and rises to 0.005 V. The input
base emitter voltage is now 0.005 V and the second base emitter voltage is
reduced by 0.005 V. The result is that more current flows in the first transistor
and less current flows in the second transistor. Because of the constant current
source, the sum of the two collector currents is a constant. The change to the
collector voltages depends on the gain of the transistors. In this example, assume
the collectors change 0.5 V. The first collector voltage will drop from 10 V to
9.5 V and the second collector voltage will rise from 10 V to 10.5 V. With this
information we can calculate the voltage gain. The input changed 0.01 V and
each collector changed 0.5 V. This is a gain of -50 to the first collector and a
gain of +50 to the second collector. The gain is associated with an offset of about
10 V.
The differential pair has no gain when both bases were raised or lowered in
potential by the same amount. A signal that is common to both inputs is called a
commonmode signal. In this circuit the commonmode gain is practically 0. In
effect, the commonmode signal is rejected. The gain to the difference signal is
called normal-mode gain.
LEARNING CIRCUIT 29
You will need (in addition to your circuit board and measuring
equipment) :
3 2N3904 transistors
Build the circuit shown in Figures 4.8 and 4.9. Note that it does not
require a matched pair of transistors to illustrate the principle. The
regulated ± 15-V power supplies are used for this circuit. The constant
current source is set for 10 mA. After you get the circuit built, make
sure that the two collector voltages are about 10 V each. Since the gain
can be 50, any input signal from a generator should be attenuated by
about 100:1 so that the signal level can be easily adjusted. A suggested
voltage divider for each base is a series 100-ku resistor and 1-kS2
resistor. This divider attenuates the signal from your function generator.
When a 1-V sine wave signal generator is applied between each 100-
ku2 resistor and common, the signal on that base is about 0.01 V. After
gain, the collector signal voltages should be about 1 V. The same signal
level should appear on each collector.
If your oscilloscope has A and B inputs, you can sum the two
collector signals. Be sure the oscilloscope inputs are ac coupled. If the
result is 0, it proves the signals are of opposite polarity or balanced. Tie
the two 100-kS2 resistors together and connect the signal generator
between these two resistors and common. Observe the collectors for
signal. This is the commonmode test. The collector signal should be
very small.
In our example of gain, the input to the first base is a signal with respect to
the zero reference conductor. This is called a single-ended signal. The two
collector signals are equal and opposite in polarity. Remember, we must ignore
the static operating voltage of 10 V. The input signal changes from 0 to +0.01 V
and the output changes ±0.5 V around the nominal operating voltage. The two
output signals are called a balanced signal. A balanced signal pair means that
one signal goes positive when the other goes negative. For a balanced signal pair
the average value is 0. Balanced signals are often used in driving long cables.
Certain kinds of noise coupling can be eliminated by using this technique.
The controlling element is called a gate. The voltage between the gate and the
source controls the current flow through the channel. The gate is actually
insulated from the source and the drain. When the gate is formed as an insulated
junction, the device is called a junction FET, or a JFET. When the gate is
insulated by a metal oxide, the device is called a MOSFET or metal oxide
semiconductor field effect transistor. There is gate current, but it is usually
nanoamperes instead of microamperes. For an n-channel FET the drain voltage
is positive with respect to the source.
There are four basic types of FET devices. The conducting channel can be n
or p material and the devices can be enhancement or depletion types. The
enhancement FETs do not conduct unless there is a gate voltage. The depletion
mode devices allow current to flow at 0 gate voltage. Some devices have both
enhancement and depletion regions of operation. For all FETs, the gate voltage
controls the electric field along the channel. It is this field configuration that
controls the flow of current in the channel. The absence of a pn or np junction
means that the voltage drop from the source to the drain can be very small.
When the transistor is turned fully on, the channel is a low resistance. When the
channel is turned off, it is a very high resistance. This makes the FET an ideal
switching device.
A FET does not have beta, as there is no controlling gate current. The voltage
curves for a typical n-channel JFET device are shown in Figure 4.9. In most
circuits the FET can be treated very much as a junction transistor. There are FET
followers, FET gain circuits, and FET pass elements in power supplies. There
are IC amplifiers made from FET devices. If you understand how transistors are
applied to a circuit, the FET should offer you no difficulty. For this reason we
limit our Learning Circuits to transistors. A problem can arise in using depletion
mode devices. The needed gate voltage may extend beyond the available power
supply voltage. Typical operating curves for a FET transistor are shown in
Figure 4.10.
There are logic circuits made from FETs. In one family of devices both p and
n channel material are used in the form of stacked emitter followers. These
components are known as CMOS, which stands for complementary metal oxide
semiconductor.
The symbols used for FETs are sometimes confusing. Manufacturers attempt
to show the construction of the device in the symbol they use for it, but these are
not always quite clear. A few of the common symbols are shown in Figure 4.11.
FET devices are particularly sensitive to handling; any slight electrostatic
discharge (ESD) can destroy the component. For this reason, FET devices are
shipped in conductive plastic to limit generating a charge when the components
rub on the plastic. An accumulated charge represents an electric field that can
cause arcing. The leads of a FET device should not be touched when handling. It
is bad practice to throw FET devices into a pile with other components and then
attempt to sort them out by hand.
When a pn junction is properly doped, the current in the junction can generate
light. Light results when electrons release energy upon returning to their normal
state from a higher energy state. The color of the light is directly related to the
energy levels associated with the two states. The light emitting diode or LED
operates in the forward direction of the diode. LEDs are not designed to
accommodate a high reverse voltage. Typical LEDs operate in the current range
5 to 20 mA. Figure 4.12 shows how a transistor can be used to turn on an LED.
The collector resistor is necessary to limit the LED current.
LEARNING CIRCUIT 30
Building a Transistor Switch
1 2N3906 transistor
1 LED diode
Phototransistors
When light strikes a semiconductor material, the absorbed light energy (photons)
can knock electrons out of their normal positions. In the base region of a
transistor, these free electrons allow transistor action. If there is a collector
voltage, then collector current results. This is the reason why semiconductor
devices are constructed in plastic in such a way as to limit any light entry.
The SCR is constructed of four layers of doped silicon. The layers are pnpn.
This arrangement of semiconducting material is known as a Shockley diode. The
outer two layers are the p and n material of a standard diode. The outer n
material is the cathode. The center np material is the SCR gate. The forward-
conducting direction of the diode is to the outer n layer. When the gate is made
positive with respect to the cathode, the SCR will conduct in the forward
direction. If the gate is at the cathode potential, the SCR will not conduct. The
symbol and the construction of an SCR are shown in Figure 4.14.
A triac consists of two SCRs in parallel. The first SCR controls the first half-
power cycle and the second SCR controls the second half-power cycle. The gate
voltage must be positive with respect to the corresponding cathode to turn the
triac on in that halfcycle. The symbol and construction of a triac are shown in
Figure 4.15. The word "triac" is a coined word implying three states. These three
states are o/fand the two directions of on.
Triacs are used in most household light dimmers. An RC delay circuit is used
to turn the triac on during each halfcycle. Changing the resistor value controls
the point in the cycle when the triac conducts. A simple version of this approach
is shown in Figure 4.16.
When the voltage between the gate and terminal 1 exceeds a threshold
voltage, the triac conducts. When the adjustable resistor (poten tiometer) is set to
0 0, the triac conducts for the entire cycle. As the resistance increases, the
capacitor delays voltage on the gate and the triac conducts later in each
halfcycle. The resistor and capacitor values must be set so that at maximum
resistance the triac never turns on. Note that during the time the triac is
conducting, the line voltage appears across the lamp and the capacitor is
discharged. There are many versions to this circuit, including a time delay using
two resistors and capacitors, bias voltages applied to the gate, and resistors or
zeners in series with the gate. These circuits are used to accommodate different
triac characteristics.
Figure 4.15 The symbol for a triac and its construction
Figure 4.16 A light dimmer circuit using a triac
Transistors as Switches
We have discussed the transistor as a component that can provide gain. When
the transistor is used at the extremes of its operation, it can serve as a switch.
When there is an excess of base current, the collector cur rent increases until
there is a small collector voltage. When the base current is 0, the collector
current is 0 and the collector voltage is at a maximum. This is the same as using
a mechanical switch. When the switch is open there is no voltage across the
resistor, and when the switch is closed the power supply voltage appears across
the resistor.
LEARNING CIRCUIT 31
1 2N3904 transistor
Build the circuit shown in Figure 4.17 on your board, using your own layout.
Connect a square wave generator set to 3 Hz to the control resistor. If the square
wave is symmetric about 0 V, then clamp the input base so that it cannot go
negative. The arrow of the diode goes to common. The LED will turn on three
times per second.
The square wave voltage on the secondary coil of the transformer can be
rectified to form a new dc power supply. If the ratio of secondary turns to
primary turns is 2:1, the new voltage will be double the original power supply
voltage.
This circuit is at the heart of many switching power supplies used in today's
electronics. The switching frequency can be 50 kHz, making the transformer a
small component. At this frequency the power supply filter capacitors can also
be fairly small.
There are several additional points regarding the circuit in Figure 4.18. On the
first cycle the core material starts out with the induction flux B = 0. The means
that the core material may saturate on the first cycle. Within a few cycles the B
field centers itself so that the field moves symmetrically between two limits. The
transformer must be designed so that the core material, the number of turns, and
the applied voltage result in a B field that never saturates the core material.
Typically, at 50 kHz the primary coil has somewhere between 10 and 20 turns.
With this small number of turns, in some cases it is possible to wind your own
transformers.
The transistors will self-destruct if they are both turned on at the same time.
The transistors will overheat even if this time period is as short as 1 its. One
technique that is used is to slow down FET transistors by using a resistor in
series with gates. This makes use of the FET transistor gate capacitance to form
a low-pass filter.
Digital Switching
The electronics we have been studying, and will continue to study throughout
this book, is known as "analog" electronics. Electronics also has a quite different
branch, called "digital" electronics, which is not part of this book. However,
analog and digital electronics are not completely separate in actuality. Every
digital device uses analog electrical signals, and there are countless areas where
the two branches of electronics must work together. It is often necessary to
convert analog signals to a digital format, and vice versa. For this reason, I have
included this brief section on digital switching, even though it is not our main
topic. Transistors used as switches often provide an interface to digital devices.
There are special IC components, made up of large numbers of transistors, that
perform these tasks. We will look at them in chapter 6.
In digital circuits, the signal is either on (logic 1) or off (logic 0.) There is no
current or voltage representing any intermediate state. There is no resistance,
capacitance, or inductance, or any of the analog concepts we have been studying
so far. However, any digital device still needs analog power to make it run. The
power supply running that digital clock is still basically the same kind of power
supply you have already studied-an analog device.
Electronic designers must provide many interfaces between analog and digital
signals. Digital designers tend to focus exclusively on their digital world, and
leave the problem of interface to analog designers. So, while we are studying
transistors as switches, it makes sense to take a look at the way transistors used
as switches can be used in these interface situations.
Digital signals are usually power supply voltages. A logic 1 might be the
positive 5 V and a logic 0, 0 V. The exact voltage is not important; a voltage of
+4.5 V would still be considered a logic 1 and a voltage of +0.5 V a logic zero.
A logic signal can be developed by turning a transistor on or off. As an example,
assume a transistor clamps a logic conductor + 5 V for a logic 1. When the
clamp is turned off, a resistor connects the conductor to 0 V, or logic 0. This is
known as a pulldown resistor.
SELF-TEST
1. A constant current source provides 5 V into 1,000 S2 and 9.992 V into 2,000
Q. What is the source impedance?
2. A constant voltage source provides 5 V into 1,000 S2 and 4.992 V into 500 Q.
What is the source impedance?
6. In problem 4, reverse the input signals. What are the two collector voltages?
7. Solve problem 5 assuming the positive input terminal is 2.55 V and the
negative input terminal is 2.50 V.
8. In problem 4, change the constant current source so that the collectors are 8 V.
What changes? What are the collector voltages when the signal voltages are
applied?
ANSWERS
1. The voltage change is 5 V. The current change is 0.004 mA. The impedance is
1.25 MU.
4. The emitter voltage without signal is -0.6 V. With 2.5 V on each input, the
emitters are at 1.9 V. With 2.5 V on one base and 2.45 V, on the second base
the input signal to be amplified is 0.05 V. The emitters move half the
distance, to 2.475 - 0.6 = 1.875 V.
5. The difference signal at the input is 0.05 V. The signal change on the first
base is +0.025 V and on the second base, -0.025 V. The first transistor
multiplies this input signal change by -40. This means the first collector
changes from 9 V to 8 V. The second collector changes from 9 V to 10 V.
The collector signals are 1 V. The difference signal between the collectors is
2 V.
6. Reverse the input signal and the output signal then reverses. The two collector
voltages are 10 and 8 V respectively.
7. The answers remain unchanged. The new signal has added commonmode
content, which is rejected.
8. The new collector voltage implies that the constant current source was
increased. The gains remain unchanged. The collector signals are still 1.0 V.
10. The voltage across the resistor is 1 V. The reactance of the capacitor at 20
kHz is 796 Q. A constant current of 1 mA (rms) by Ohm's law implies a
voltage of 7.96 V (rms).
Objectives
• the concept of feedback, and how it is used to obtain positive and negative gain
A typical IC has a gain from the input to the output that is often over a million
at dc. An input attenuator of a million to one would not be a good way to use this
huge amount of gain. It must be dealt with by imbedding the IC in a circuit that
regulates the gain. The method used for this purpose is known as feedback.
Negative Feedback
The IC Package
Before we can proceed with our study of IC amplifiers, you need to become
fainiliar with the notation used in schematics for circuits that include them. The
symbol for a linear integrated circuit amplifier is a triangle, as shown in Figure
5.1.
The two input terminals are labeled + and -. The power supply voltage
connections are usually not shown on this symbol. The output terminal is usually
taken from the right at the tip of the triangle. The internal components that make
up the IC are connected between the plus and minus power supply voltages. The
common power supply conductor is not connected to the IC amplifier. This
conductor is used in the external feedback circuit that supports the IC amplifier.
LEARNING CIRCUIT 32
Assume that the IC amplifier has an openloop gain of 1,000,000 at dc. In this
application the negative input terminal is connected to the output terminal. Next
we connect the positive input terminal to the power supply common through a
resistor of 10 M. We do this because the input terminals to any amplifier must
not be left open and undefined. This circuit arrangement is a positive "gain 1"
amplifier. The gain of 1 is measured from the positive input terminal to the
output terminal. If the positive input signal is set to +2 V, the negative input
terminal and the output terminal must rise to 2 V less 2 .tV. The circuit is in
balance because the 2 tV of difference signal is the right signal to be multiplied
by 1,000,000 for an output of 2 V. To be accurate, the gain is actually 0.999999.
If the input is set to -2 V, the output terminal will adjust to -2.0 V plus 2 µV, or
-1.999998 V
Figure 5.2 A gain 1 feedback amplifier
Figure 5.3 The construction of the circuit in Figure 5.2
The circuit in Figure 5.2 has 100% negative feedback. The positive input
terminal to the IC is the input terminal to our circuit. This configuration is called
potentiometric feedback. To see how this circuit operates dynamically, place a
+2-V signal on the positive input terminal. The output starts to rise. As the
output nears 2 V, the negative input nears the value of the output. Remember,
this amplifier reacts only to differences. If the output were exactly 2 V, the input
signal difference would be 0. This implies that the output is also 0. In fact, the
output can only rise to a final value of 1.999998 V. Obviously, this is very close
to the right answer.
The output voltage of this gain 1 amplifier will follow the input voltage as
long as the voltage is within the performance specifications of the amplifier. If
the voltage is a square wave or sine wave, the output will follow. The circuit acts
like a very accurate emitter follower. The output of this IC amplifier will supply
up to ± 10 mA and ± 10 V.
The input impedance to this circuit is very high. When the input base changes
1 V in potential, the emitter also changes 1 V to within a microvolt. The input
signal current hardly changes when the input voltage changes. This is a high-
input impedance circuit by our definition. There must be base current, but this
does not count as signal current.
LEARNING CIRCUIT 33
You will need (in addition to your circuit board with mounted socket
and your measuring equipment):
1 LF353 IC amplifier
1. Add the circuit shown in Figures 5.2 and 5.3 to your circuit board.
This circuit has a gain of 1, a high input impedance, and a low output
impedance. These are the characteristics of a stacked emitter follower. This
circuit is better than a stacked emitter follower; it is less expensive, smaller, and
more accurate. This assumes that the output current and voltage are adequate.
The circuit in Figure 5.2 can be modified to have any gain from 1 to about 100.
Higher gains are possible depending on the type of IC amplifier being used. For
the moment, a gain of 10 will demonstrate how the feedback circuit works.
In the circuit of Figure 5.4, an output voltage divider attenuates the output
signal by a factor of 10. This voltage attenuator is called a feedback divider.
Figure 5.4 A positive gain 10 circuit using an
IC amplifier
Figure 5.5 The construction of the circuit in Figure 5.4
When the circuit processes a signal, the output signal voltage must adjust
until the two base voltages are equal. If one-tenth of the output is fed back to the
negative input base, the output voltage must increase by a factor of 10 to cause
equal signals to appear on the two bases. If the attenuation factor is 20:1, the
gain will be 20. The rule to calculate this is as follows: Place an input signal on
the input base. Calculate the output level that is required so that this same
voltage will appear on the second or feedback base. The gain will be the
reciprocal of the output attenuation.
The higher the gain, the larger the gain error. As an example, if the gain is 1,
and the input signal is 1 V, the error signal is 1 µV. The ratio of input signal to
error signal is 1,000,000:1. Assume the output signal is still 1 V. If the gain is
20, the input signal is /,, the previous value, or 0.05 V. The difference signal on
the differential stage must still be 1 .tV. Remember, the open-circuit gain is still
1,000,000. The ratio of input signal to error signal is now 50,000:1. This means
that the signal gain error is one part in 50,000. This is still a small error, and it
can be neglected. As you can see, excess gain (feedback factor) provides
accuracy. By requiring a gain of 20, the excess gain has been reduced by a factor
of 20 and the circuit is not as accurate. In a practical sense, the accuracy is
limited by the resistor values and not by the feedback factor. This type of
feedback circuit cannot have a gain of less than +1. Of course, an input
attenuator can always be added to limit the gain, but this solution has drawbacks,
as the input impedance must be reduced to form a practical voltage divider.
LEARNING CIRCUIT 34
You will need (in addition to your circuit board and measuring
equipment):
1 LF353 IC amplifier
Add the output attenuator in Figure 5.4 to the circuit you built in
Learning Circuit 33. Show that the gain is now +10. If your signal
generator does not have an output level adjust, it may be necessary to
reduce the generator output by using an attenuator. A series 10,000-a
and a 100-Q resistor can serve as a voltage divider to attenuate the
generator signal for our lessons. The reduced input signal is taken
across the 100-52 resistor.
The gain of the circuit in Figures 5.4 and 5.5 can be changed to
about 20 by paralleling the 1,000-52 resistor with a second 1,000-a
resistor. Observe this gain change using your oscilloscope. Be sure the
output signal at a gain of 20 is kept less than 10 V peak. You may have
to reduce the level of the input signal.
Gain is the ratio of change between an output signal and an input signal.
Negative gain means that the output change is in the opposite direction. For
example, if the output changes 2 V for a 1-mV input change, the gain is +2,000.
If the output changes -2 V for an input change of 1 mV, the gain is -2,000.
Where does this current go? There is no path into the base of the input transistor,
so it must flow in resistor R2. In fact, the output of the amplifier will adjust so
that the input voltage is essentially 0. The output voltage is equal to the input
current times the feedback resistor with a minus sign. This means that:
But I equals VIN/RI. Therefore:
The voltage gain of the amplifier with feedback is the negative ratio of
resistors, or -R2/R,. If RI = 10 kSZ and R2 = 100 kS2, the gain will be -10.
The voltage at the negative input terminal of the IC is the error signal. The
negative input terminal to the IC is the summing point. If the gain is 1,000,000,
this signal is always less than 10.tV. If the gain is 1, the largest input signal is 10
V. The error signal at the summing point is 10 gV. If the gain is 100, the largest
input signal is 0.1 V and the error signal is still 10 µV. For this smaller signal,
this error is one part in 10,000. Just like the circuit using potentiometric
feedback, the percentage error depends on the excess gain or feedback factor.
For all practical purposes, the summing point does not move. This is why it is
called a virtual ground. Later we will see why the error voltage increases at
higher frequencies. You may try to look at the summing point with your
ocilloscope, but you will be disappointed. You will add noise to the circuit, and
the signal will be too small to observe.
LEARNING CIRCUIT 35
You will need (in addition to your circuit board and measuring
equipment) :
1 LF353 IC amplifier
1 1.5-V battery
1. Modify the circuit board using the feedback circuit shown in Figures
5.6 and 5.7. Set R2 = 100 kS and RI = 10 kQ. Connect a sine wave
signal at 0.5 V 1 kHz between the input terminal and the signal
common. Adjust this signal level until the output signal of the
amplifier is 5 V. Now change the signal to a square wave and note
the same gain. Use a load resistor of 1 W.
In potentiometric feedback, the excess gain kept the two input bases at the
same signal potential. In the operational feedback circuit, the same thing is true
except that one of the bases is connected to 0 V. The excess gain requires that
the negative input terminal stay near 0 V.
Feedback circuits are self-adjusting. The signal voltages in the circuit adjust until
there is a balance at the error point or summing point. As an example, suppose a
load resistor demands more output current. This greater current must flow in the
internal resistance of the output circuit. This requires a larger drive signal. In a
feedback circuit, the error signal increases by a small amount and supplies this
extra signal. The result is that the output voltage does not appear to change. A
signal source that does not change when there is a change in output current has a
low output impedance. For most IC amplifier circuits, the output impedance is in
the milliohm range. This is a very low value that is often hard to measure.
The output signal will be a clean sine wave. The signal internal to the
feedback structure can be nonlinear. The signal driving the emitter followers will
exactly compensate for the diode problem. The output of the IC amplifier will
have a steep waveform around the zero crossings that exactly corrects for the
inability of the output stage to conduct current.
This circuit has the advantage of limiting power dissipation. When there is no
signal, there is no current in the stacked emitter portion of the circuit. When
there is a signal, the circuit will deliver this signal to the load without significant
distortion.
Figure 5.9 The distorted voltage waveform for a stacked
emitter follower
Figure 5.10 A feedback circuit using the stacked emitter
followers as an output circuit
The way feedback corrects for internal distortion can be shown by modifying the
circuit shown in Figure 5.2. Normally the feedback is taken from the output
terminal of the IC amplifier. As an experiment, we are going to intentionally
distort the gain by placing back-to-back parallel signal diodes in series ahead of
the load resistor. This circuit is shown in Figures 5.11 and 5.12. A distorted sine
wave with a horizontal flat segment around each zero crossing occurs at terminal
B. If the feedback is taken from the load resistor with the diodes in place, the
distortion is removed. The diodes are now inside the feedback loop.
Figure 5.11 The signal patterns in a feedback circuit with added diodes
LEARNING CIRCUIT 36
Your will need (in addition to your circuit board and measuring equipment) :
1 LF353 IC amplifier
(Continued)
3. Set the signal to 2 V peak at 1 kHz. Observe the signal at the output of the
amplifier ahead of the diodes and after the diodes.
5. Observe the signal at points A and B. The signal at A will have a very steep
transition around the zero crossings. This is the error signal multiplied by the
gain of the amplifier.
Figure 5.12 The construction of the circuit in Figure 5.11
When we studied passive circuits, we used a pointer system to show the effect
capacitance had in delaying a sine wave. We also saw that sine waves that are
shifted in phase do not add and subtract directly. All circuits have small
capacitances that limit their amplitude response as a function of frequency. This
means that phase shift and delay are a part of every circuit design. A feedback
signal cannot subtract from an input signal unless the two signals are in phase.
Even a few degrees of phase shift creates a problem.
To control and limit the phase shift, the openloop gain of the amplifier must
begin to lose gain at a very low frequency. As an example, suppose the gain is
1,000,000 as it starts down at 10 Hz. The gain will be 100,000 at 100 Hz and
10,000 at 1 kHz. To maintain this same slope, the gain reaches unity at 10 MHz.
This does not mean that the amplifier can handle 10-MHz sine waves. It only
means that the openloop phase character is controlled to this frequency. An
amplifier with this controlled phase shift is called an internally compensated
amplifier. Fortunately for the user, this difficult problem has been solved by the
IC designer.
To understand positive feedback, imagine that if your car was going faster
than you wanted, you responded by increasing rather than decreasing its speed.
The car would then go faster and faster until it broke down or ran out of gas. If
the car was going slower than you wanted, positive feedback would be required
to further decrease the speed until the car stopped. In electronic circuits, positive
feedback generally results in an oscillation. The circuit goes in one direction
until it can't go any farther, and there is a sharp change to the opposite direction-
this is the oscillation. In circuits with too much phase shift, there may be no
oscillation, but the circuit might border on instability. An unstable response is
shown in Figure 5.13.
Instability problems are most apt to occur when the bandwidth and the
feedback factors are high. If there is an instability, the oscillation can be low-
amplitude at a frequency above 10 MHz, which might go unnoticed. It is always
a good idea to test a circuit for stability using a square wave voltage. The circuit
should be able to accept a capacitance load and not show signs of oscillation. To
check for stability, it is a good idea to try several values of capacitance loading
from 100 pF to 0.01 .tE When a circuit is close to being unstable, there is
"ringing" (the oscillation shown in Figure 5.13) on the leading edge of the output
square wave. Designs with ringing are undesirable. A change in temperature or a
change in load could take the circuit over the edge.
In the circuits we have discussed so far, the power supplies have been
symmetrical about a midpoint. This midpoint has been associated with the input
circuit common and with the feedback design. Remember, this midpoint is not
connected to the IC amplifier. In effect, the amplifier operates from a single
potential difference.
One way to solve this problem is to use a voltage divider on the positive input
to the IC amplifier. Feedback forces the negative input to be near this same
potential. The voltage divider establishes a common or zero for the circuit.
If one side of the power supply is grounded, the voltage divider can still be
used. The potential at the positive input is the reference potential for the
amplifier. In effect, the input and output signals are referenced to this potential.
If the signals originate referenced to the power supply common, then a coupling
capacitor can be used to block any offset. A typical circuit is shown in 5.14.
Figure5.14 An amplifier circuit using a single supply voltage
A differential amplifier can limit this current flow and at the same time reject
the difference in potential between reference conductors. This circuit is shown in
Figures 5.15 and 5.16.
This circuit has four feedback resistors. The resistor values are selected so
that R,/R2 = R3/R4. In the circuit analysis we will use, all the signal voltages are
referenced to the power supply common. The gain to B is R2/R,. This requires
the signal A to be 0 V. The gain to A is -R2/R,. If A = B, then the two signals
cancel and the output signal is 0. There is only gain to the difference between A
and B. This gain is the ratio of R2/R1. The amplifier responds to the difference
signal and ignores the average signal. This is the meaning of the term differential
amplifier. If the resistors are matched to 1% resistors, the average value will be
rejected to within 1%. The rejection of the average value is called commonmode
rejection.
LEARNING CIRCUIT 37
You will need (in addition to your circuit board and measuring equipment):
1 LF353 IC amplifier
1. Build the circuit shown in Figures 5.15 and 5.16. Use R2 = R., = 10 kS2 and
R, = R3 = 1 W.
2. Build a voltage divider across the power supply to obtain several voltages.
The resistors can be 10 kS2 and four 330 Q. The voltage across each 330-52
resistor is about 0.437 V.
3. Connect the IC amplifier input terminals across any one of the 330-a
resistors. The gain is 10, so the output voltage should be 4.4 V. You should
get the same result regardless of which resistor you select.
4. Reverse the leads and notice that the output voltage reverses polarity. Leave
the input terminals connected to one of the bottom three 330-Q resistors.
5. Place the signal generator with a 1-V sine wave at 100 Hz across the top 330-
Q resistor. Notice that this signal does not appear in the output. This signal is
added to both inputs and is rejected.
6. Turn off the power supply. The gain of the amplifier is 10.
10. Place the input leads across one of the power supply diodes. Observe the
signal at the output of the amplifier. This signal is the voltage across the
diode attenuated by a factor of 10, referenced to the common conductor of the
power supply. Do not attempt to measure the diode voltage by placing the
oscilloscope across the diode. The generator and the oscilloscope can both be
safely grounded, and this will short out the transformer secondary winding.
This ability to look at signals anywhere in a circuit is a valuable asset.
This circuit can be used to measure the voltage across any circuit element,
including the diodes in the rectifier circuit. Here the commonmode voltage can
be 50 V peak. To use the amplifier for this application, the gain must be less than
1. If R2/R1 = 0.10, the differential amplifier can accommodate this 50-V signal.
The circuit in Figures 5.17 and 5.18 provides an attenuation slope that is
proportional to the square of frequency. Stated another way, the attenuation
slope is 40 dB per decade. At 10 times the frequency, the attenuation factor is
100. This is called a second-order filter.
LEARNING CIRCUIT 38
You will need (in addition to your circuit board and measuring
equipment) :
1 LF353 IC amplifier
2 10-ku resistors
2. Set the two resistors equal to 10 kS2 and the capacitors equal to C2 =
0.02 tF and C, = 0.01 µE Check the frequency response using sine
waves. Note the -3-dB point, the frequency where the attenuation
factor is 0.1, and the frequency where the attenuation factor is 0.01.
3. Double the value of C2 and notice that the frequency response has a
significant peak. Also notice that the overshoot for a square wave has
increased.
This filter has the same characteristics as the RLC circuit described in Figure
2.9. This is an example of where an active circuit has the same response as a
passive circuit using an inductor. However, this circuit can be made to operate at
frequencies as low as 1 Hz, which is almost impossible with inductors. Inductors
saturate, and they have their own natural frequency. They are bulky and
expensive. The circuit approach has the further advantage of providing a low
output impedance.
The circuit can be designed to have different cutoff frequencies and different
amplitude responses near the cutoff frequency. The exact -3-dB frequency
depends on the ratios between capacitors and resistors. A good approach in
design is to set the two resistors equal. When C2/ C, = 2 and R, = R2, the
frequency response is near optimum and the square wave response will have a
7% overshoot. If C2/C1 = 1, the square-wave response will have no overshoot.
The -3-dB point will be approximately
When the capacitors and resistors in Figure 5.17 are reversed, a highpass filter is
formed. The filter is optimum when the capacitors are equal and the resistors
have a ratio of 2:1. The terminal slope for this filter is 40 dB per decade. This
filter is shown in Figures 5.19 and 5.20.
(Continued)
LEARNING CIRCUIT 39
1 LF353 IC amplifier
2 0.01-µF capacitors
3. Test the circuit using a square wave signal. Note the response
waveform. When the input waveform goes positive, the output
follows and immediately returns to 0 overshooting 0. The same thing
happens when the input wave goes negative. These undershoots and
overshoots are characteristic of highpass filters. A highpass filter
cannot pass dc, so the average voltage in the output must be 0.
SELF-TEST
4. An operational feedback circuit has an input resistor of 200 kc2. The feedback
resistor is 560 kS2. What is the gain? Indicate the gain polarity.
6. An IC amplifier has an openloop gain of 100,000 that starts losing gain at 100
Hz. The closedloop gain is 20. What is the feedback factor at 10 kHz?
10. An output emitter follower circuit has 3% distortion. The feedback factor at
1 kHz is 200. What is the expected distortion in the output waveform?
11. An operational feedback circuit has an input 100-ku resistor. If the maximum
output voltage is 10 V and the gain is 10, what is the maximum input current?
12. A differential circuit uses four equal 20-ku2 resistors. What is the gain to the
difference signal? What is the commonmode gain?
13. A low-pass second-order filter uses 20-kc resistors. What are the capacitor
values if the -3-dB frequency response is to be at 2 kHz?
14. In problem 13, what are the capacitor values if the -3-dB frequency is to be
20 kHz?
15. A highpass second-order filter uses 0.1-µF capacitors. What are the resistor
values if the -3-dB frequency is to be 20 Hz?
ANSWERS
1. 10 V/200,000 = 50 µV.
2. The attenuation is 1,000/4,000 and the gain is 4. For reversed resistors, the
gain is 4/3.
3. 1.0.
5. Assume the input is 0.1 V. The output voltage is 3 V. The voltage at the
junction of the feedback attenuator is -0.3 V. The voltage across the top
resistor is 2.9 V. The current in the 200-i2 resistor is 0.1/200 = 0.5 mA. The
current in the 50-kQ resistor is 0.1/50,000 = 2 I.A. The total current in the top
resistor is 502 µA. By Ohm's law, the resistor is 2.9/0.000502 = 5.776 W.
6. The ratio of 10 kHz to 100 Hz is 100. The openloop gain is only 1,000. If the
closedloop gain is 20, the feedback factor is 50.
10. 3%/200=0.015%.
11. The maximum input voltage is 1.O V. The input resistance is 100 k12. The
maximum input current is 10 µA.
15. The nominal resistor values are 79.6 W. To allow the ratio of resistors to be
2:1, multiply R2 by V and multiply R, by 1/N/-2. The result is R, = 56.25 kit
and R2 = 112.58 W.
Objectives
Integrated circuits are used to perform many tasks besides the one we looked
at in chapter 5, providing voltage gain. They are used in waveform generation,
timing circuits, voltage regulators, signal measurement, oscillators, comparator
circuits, switching, and buffering, to mention just a few applications. Units with
highfrequency performance are used to process signals in radio and television. In
fact, the majority of today's electronic designs make use of integrated circuits. In
many mass-produced products, several circuit functions are even combined into
one IC. In this chapter we will examine several of these applications in more
detail.
Voltage Regulators
To learn about voltage regulators, you are going to build one. A typical
positive 5-V regulator circuit is shown in Figure 6.1. This circuit is far more
stable than the emitter follower circuits you built in chapter 4. The output
voltage adjusts so that the signals on the inputs of the IC are equal. The voltage
divider determines the regulated voltage. A negative power supply is
unnecessary because all of the operating voltages of the IC are positive. This
circuit should always have an electrolytic capacitor connected from the regulated
output to the common. A typical value might be 10 tE This is also required for a
threeterminal regulator.
LEARNING CIRCUIT 40
You will need (in addition to your circuit board and measuring equipment):
1 LF353 IC amplifier
1 TIP29A transistor
(Continued)
The current sensing resistor is R1. The unity gain differential stage measures
the voltage drop across this resistor and feeds this signal back to the negative
input. This is an example in which the feedback path has an active circuit. This
circuit will overload without an output load resistor. If the sensing resistor is
100Q and the input voltage is 1 V, the output current will be 10 mA. If the load
resistor is 500 S2, the output voltage will be -5 V. The output voltage is -6 V if
the load resistor is 600 Q. This circuit provides a current source for ac and dc
signals.
A constant current source that is often used in industrial measurement is
shown in Figure 6.4. A sensor that measures pressure or flow has an internal
current amplifier. The load resistor for the output of this amplifier is placed at
the end of the signal loop, which might be hundreds of feet away.
The load current must flow in resistor RFB. This is the current sensing
resistor. The feedback circuit requires that the voltage across RFB equal the
input voltage. This means that LOUT = VIN/RFB. The voltage across the load
resistor changes with resistance value. As an example, assume that RFB = 100
Q. If the input voltage is 1 V, the current in the output is 10 mA. A load resistor
of 500 S2 will have 5 V across its terminals. If the load resistor is 400 S2, the
voltage is only 4 V. The voltage drop across the connecting wires is ignored.
In an operational feedback circuit, the summing point does not move. If the
input voltage is constant, the current to the summing point is also constant. If the
feedback element is a capacitor, this constant current must flow into the
capacitor. The result is a steady increase in voltage across the capacitor. This
voltage is also the output voltage. The output voltage is the integral of the input
voltage. This integrator circuit is shown in Figure 6.5.
The output of the integrator circuit inverts the sign of the integral. To see the
uninverted integral, a second unity gain inverting amplifier can be placed after
the integrator. It is helpful to see graphically what the ideal integral is for various
input voltage waveforms. At t = 0, the capacitor is assumed to have no charge on
its plates. The charge is added or subtracted by the feedback amplifier. These
integrals are shown in Figure 6.6.
Consider an input voltage that is 1 V for 0.1 second. Assume the output
voltage is a ramp that goes minus to -1.0 V. If the input voltage then returns to 0
V, the output must stay at -1 V Will the integrator hold the voltage for a minute
or an hour? In practice, an IC integrator will begin to drift away from its proper
value after a few seconds.
Figure 6.6 The integration of various waveforms
An explanation that provides some insight into the problem involves the
circuit response at low frequencies. Assume the closedloop gain at 100 Hz is 10
and the reactance of the capacitor is 100 ki2. At 10 Hz the reactance is 1 MS2
and the closedloop gain of the circuit is 100. Consider a frequency as low as 0.1
Hz. The closedloop gain is 10,000 and the reactance is 100 M. It is easy to see
that as the frequency gets lower and lower, the closedloop gain will eventually
reach a limit. It cannot exceed the openloop gain, and this is the point where
there is no feedback. In other words, the integrator is no longer functional.
Another thing happens: there is input base current. This current is no different
from input signal current except it is much smaller. The result is an error in the
integral that grows with time.
The integrator in Figure 6.5 is a valuable tool, but its use has a time
constraint. Before the error becomes too large, the charge on the capacitor must
be removed and the integration must start over. A digital up-down counter is also
an integrator. The last count can be held in a register until the computer is turned
off. This is one advantage a digital integrator has over an analog integrator.
The Comparator
An IC amplifier responds to the difference signal at its two inputs. Suppose the
negative input is held at 4.5 V. If the positive input is greater than 4.5 V, the
output is positive. If the positive input is less than 4.5 V, the output is negative.
If there is no feedback and there is a great deal of openloop gain, the output will
respond with a full-scale signal for a very slight change in the input signal at
right around 4.5 V. The IC amplifier is functioning as a comparator. This circuit
is shown in Figure 6.7.
LEARNING CIRCUIT 41
1 LF353 IC amplifier
1 1N4148 diode
1 0.01-µF capacitor
The integrator circuit can be used to provide a time delay. The circuit in Figure
6.11 has a short across the charging capacitor.
When the switch opens, the voltage across the capacitor begins to ramp up.
When this voltage reaches the comparator reference voltage level, the
comparator output goes positive and can operate an LED, a relay, or another
transistor. The integrator resistor and capacitor control the delay time. A time
delay is sometimes required for power-up, to limit access time, operate a
temporary light, or control exposure time in photography. With the right IC
amplifier, if the capacitor leakage current is low and the capacitance is high
enough in value, the time delay can extend to several minutes.
Figure 6.11 A time-delay circuit
Trigger Circuit
When a slowly changing signal hits a critical level, it is often desirable to make a
positive decision to do something. The situation might be the overheating of a
bearing. A trigger circuit responds to the critical temperature level and
transitions from 0 to the plus power supply voltage. This voltage can then
operate a relay to sound an alarm. The output will not transition back to 0 unless
there is a significant reduction in the input signal. For example, in Figure 6.12
the negative input voltage is set to 5 V.
If the signal voltage reaches 5 V, the trigger circuit output goes from 0 to 15
V. If the signal falls below 3 V, the trigger circuit output transitions back to 0.
The gap between 5 V and 3 V is called hysteresis. This circuit is known as a
Schmidt trigger.
After the threshold is reached, the positive feedback from the output increases
the input signal to 7 V. The circuit cannot reset itself until the input voltage
drops back to 3 V. This hysteresis stops the trigger circuit from oscillating
around the transition point. If there is oscillation the trigger voltage would be
unusable.
The Multivibrator
A class of circuit can be made from IC amplifiers or transistors that have two
operating states. As an example, an output voltage could be stable at 0 V and
+15 V These two voltages are often near the limits of the power supply.
Multivibrator circuits that change state in response to a pulse or an input signal
are called bistable multivibrators or "flip-flops." Circuits that have voltages that
constantly transition back and forth between two end states are known as astable
multivibrators. They are a class of oscillator.
Circuits that have two stable states are used extensively in digital logic. It is
worth noting that digital ICs are made from analog circuits that include gain,
output drivers, comparator clamps, zeners, and diodes. In the early days of
digital electronics, all of the logic functions were built from discreet transistor
circuits. These circuits are still useful, and it is helpful to know how they
operate.
In the following sections we will discuss two circuits, the bistable and the
astable multivibrator, which are analog circuits used in digital design. A
multivibrator circuit can be recognized by its structure. It usually consists of two
gain elements that are cross-coupled-gain element A drives gain element B and
gain element B drives gain element A. The coupling is such that one of the gain
elements conducts while the other is turned off. In the bistable multivibrator an
input pulse will cause the gain elements to change roles. In the astable
multivibrator the roles change automatically. After a transition a capacitor
discharges, allowing the next transition.
A positive pulse at the input terminals couples to both bases. The pulse cannot
add to the current in Ql, because the transistor is already fully conducting. The
pulse can, however, cause Q2 to conduct. The collector voltage of Q2 drops. The
voltage divider to the base of Ql couples this voltage drop and reduces the
current flow in Q1. The rise in the collector voltage of Ql further increases the
base voltage of Q2. The end result is that transistor Qi is turned off and transistor
Q2 is turned on. From symmetry, a second input pulse causes Ql to conduct and
Q2 to be turned off. In this way the circuit can be made to flip-flop back and
forth.
LEARNING CIRCUIT 42
You will need (in addition to your circuit board and measuring equipment) :
2 2N3904 transistors
2 1N4148 diodes
1 0.001-µF capacitor
1 0.001-µF capacitor
2. Measure the voltage on the two collectors and determine which transistor is
turned on.
3. To change the state of the mutivibrator, touch point C to the plus 15-V. This
generates a pulse to change the state of the multivibrator.
4. Touch point C to the power supply a second time. The circuit should again
change state. Verify this by observing the collector voltages.
Figure6.15 An astable multivibrator
The moment Q1 starts to conduct, its collector voltage falls to near zero. The
base voltage on Q2 also falls to a negative value. This is because Cl and R1
make a highpass filter and the leading edge comes straight through this filter. As
capacitor Cl charges through R1, the base of Q2 rises in voltage. When the base
of Q2 is positive with respect to the emitter of Q2, it starts to conduct. The
collector voltage on Q2 starts to fall, and this turns off transistor Q1. This double
action causes the state of the mutivibrator to switch. During the time C1 is
charging, C2 is discharging, and during the time C2 is charging, C1 is
discharging. The RC time constants determine the frequency of oscillation.
Crystal Oscillators
A crystal is a thin layer of quartz. Quartz has the property that a voltage is
developed between two surfaces when pressure is applied. If a voltage is placed
across the conductive surfaces, a strain results. If the surfaces of the crystal are
plated, then the voltage resulting from a strain can be easily sensed with a high-
input impedance voltmeter. A wafer of quartz with conductive surfaces has all
the characteristics of a capacitor.
The quartz crystal is a mechanical structure that can store potential energy.
The potential energy is stored in stress. This potential energy can transfer to
kinetic energy when the crystal moves. All the elements of an oscillator are
present. The crystal mass stores potential and kinetic energy. If energy can be
supplied per cycle, the mechanical vibration can be sustained. The result is a
sinusoidal voltage across the plates of the crystal. The crystal oscillator
frequency is very stable, and very little energy must be supplied per cycle to
sustain an oscillation. A typical circuit is shown in Figure 6.16.
Figure 6.16 A crystal oscillator
Counting digitally requires eliminating all the number symbols except 0 and
1. The first sixteen numbers are 0, 1, 10, 11, 100, 101, 110, 111, 1000, 1001,
1010, 1011, 1100, 1101, 1110, and 1111. Note that one of the numbers is a zero.
Nothing is changed if we add leading zeros, so the sequence can also be written
as 0000, 0001, 0010, 0011, 0100, 0101, 0110, 0111, 1000, etc. In a digital
circuit, bits are represented by the states of transistors. A bit is a single logic
state, a 0 or a 1. For example, if the transistor is turned on, it might stand for a 1,
and if it is turned off, it might stand for 0.
Each digital position is called a bit. In this example, we have a fourbit number
and we have counted from 0 to 15. If the counting had continued to include 8
bits, then we would be able to count to 255. Eight bits are called a byte. A string
of 1s and Os is called a digital word. A single bit could be 5 or 10 V, depending
on the type of logic involved. The voltage itself has no meaning. The presence of
a voltage represents a logic 1 and the absence of voltage represents a logic 0.
In order to represent voltages less than 6.4 V more accurately, more bits are
required. If 8-bit words are used to cover the range 0 to 6.4 V, then an increment
of 1 bit represents a change in voltage equal to 0.025 V.
A very useful resistor network provides the basis for converting digital signals to
analog signals. This network in Figure 6.17 is called an R-2R ladder. There are
two values of resistor. The accuracy of the resistors is not as important as the
ratio between them.
Figure 6.17 An R-2R ladder network
Now assume all of the voltages are equal. This ladder circuit allows us to sum
a group of currents from a fixed voltage source and attenuate these currents by
factors of two. If the ladder has eight entry points, then the current that flows can
be increased in steps where the largest step is 1 and the smallest step is 1/256. As
an example, if the first current is 1 mA, the second current is / mA, and the third
current is 4 mA. Our problem is to control which voltages are to be connected.
The maximum current in this arrangement is 1.996 mA. The R-2R ladder will
only work if the voltage sources are set to 0 V when they are not used. A voltage
source of 0 V is a short circuit.
Circuits that convert digital signals to analog signals are called D/A converters.
The conversion can be accomplished using an R-2R network. The currents that
flow in this network can be controlled by a set of transistor switches. The
switches connect the 2R resistors to either a reference voltage or to 0 V (circuit
common). The switches are controlled directly by the logic levels of the digital
word. When the logic is a 1, the switch is turned on and a weighted current flows
into the summing point. The amplifier responds to the sum of the network
currents with a negative voltage that represents the digital word. A fourbit
version of this D/A is shown in Figure 6.18.
The full-scale voltage for the D/A converter can be set by the reference voltage
or by the gain of the operational amplifier. If a count of 15 equals 10 V, then
each count represents 0.666 V.
Figure 6.18 A fourbit D/A converter
AID Converters
There are several techniques available for converting an analog signal into a
digital word. Accurate conversions at a rate exceeding 200,000 samples per
second are practical. At these high speeds there is little time available to take a
series of logical steps. This makes a high-speed A/D converter a very
sophisticated circuit.
A very simple way to determine the digital value for an analog signal is to
advance a digital counter through its set of values. A digital counter is a circuit
that stores a set of bits in a circuit called a register. An external pulse advances
the binary number stored in the register. The bits in the counter are converted to
an analog signal (D/A converter) by using an R-2R network and an operational
amplifier. This D/A converter output is compared with the sampled voltage.
When a comparison is reached, the counter is stopped and the stored digital
count is the correct digital word. There is always a round-off error, as there are
only so many values available, depending on the number of bits.
Another type of A/D converter is called a pipeline flash converter. The sampled
signal is applied in parallel to a group of comparators. This results in an
approximate measure of the signal. The output of these comparators is connected
to a D/A converter. The output of this D/A converter is subtracted from the
initial signal. The difference signal is amplified and again sent to the
comparators. The output of the comparators is again connected to the D/A
converter. The output of the D/A is again subtracted from the input. This process
is repeated until all of the bits in the A/D converter have been determined. This
pipeline approach takes more time than the true flash system, but it provides for
a great deal of accuracy.
One, two, or four arms of the bridge can be active. An active arm changes
resistance when the parameter of interest changes. If one arm is active, then the
other three arms are usually passive resistors. The resistors that make up the
passive arms must not change value when the temperature changes.
For stress and strain measurements, the active arms are made from a long strip of
thin resistance wire that is fan-folded and bonded to a plastic carrier. This active
arm is called a strain gauge. It is oriented and bonded to a structure so that if the
structure changes dimension, the gauge wire is stretched or compressed. The
change in dimension changes the resistance of the strain gauge. If this gauge
element is one arm of a Wheatstone bridge, a difference signal is generated
across the signal terminals of the bridge. Strain gauge elements can have a
resistance in the range 50 to 300 Q. Under stress, these resistors might change 1
or 2%.
Figure 6.19 A typical Wheatstone bridge circuit
The signal that is developed across a bridge is greatest when all four arms are
active. Two gauges must increase in resistance and two must decrease in
resistance to produce an optimum signal. If all gages increase in resistance by
the same amount, the gauge remains balanced.
SELF-TEST
1. The current source in Figure 6.2 has a 200-0 sensing resistor. The input
voltage is 3 V at 1 kHz. What is the output current? What is the output
voltage if the load resistor is 500 S2?
2. In problem 1, what is the output voltage if the load resistor is 400 S2?
3. A current source in Figure 6.2 has a 400-52 sensing resistor. The input voltage
is -4 V at 3 kHz. What is the output current? What is the output voltage if the
load resistor is 400 S2 shunted by 0.1 µF?
5. A 20-Hz 10-V square wave signal is placed into the integrator of problem 4.
What is the output waveform? What is the peak-to-peak output voltage?
10. An 8-bit digital word represents voltages from -10V to +10 V. What is the
digital value for 10 V?
11. In problem 10, if all Os represent -10 V, what does the digital value 1000
0000 represent?
12. In problem 10, what is the voltage difference for the least significant bit?
Hint: The count maximum is 28 = 256.
13. Can the first bit be used to tell the sign of the voltage?
14. Can you find a way to use a digital representation of a voltage and separate
the sign of the voltage from its magnitude of the voltage?
ANSWERS
2. 6V.
3. The current is 10 mA. The impedance at 3 kHz is 267 Q. The voltage is 2.7 V.
6. The voltage is onequarter the percentage resistor change times the excitation,
or 62V2 mV.
7. The gain is 160.
11. o v.
14. To obtain the magnitude of the negative number, swap 'Is for Os and add 1.
This is called the 2s complement.
Objectives
• some techniques for laying out circuits to avoid noise and interference
We have now covered many basic circuits and seen how they operate. Of
course, we have only been able to touch upon our subject, and have not been
able to go into details of constructing these circuits. If you continue your study
of electronics you will come to appreciate that in many applications performance
depends on the exact arrangement of components. In other words, the
component geometry is an important part of the design.
The circuits we have studied can be analyzed using Ohm's law, plus an
understanding of transistors and integrated circuits, the components that allow us
to amplify, rectify, control, and monitor various electrical signals. However, in
many situations, circuits include structures that are not under our control. The
earth, for example, is a conductor that is shared by all users. It has resistance, but
it is not a resistor. The conduit in a building or residence may be grounded (tied
to earth) along with our electronics. The interconnection of equipment grounds is
not a part of any schematic.
The next chapter discusses the basics of electricity. You will find that voltage
is defined in terms of the electric field. Another field, the magnetic field, exists
whenever there is current. Capacitance is the ability to store electric field energy,
and inductance is the ability to store magnetic field energy. Whenever there is a
voltage there is an electric field, and whenever there is a current there is a
magnetic field. A changing electric field in a capacitor is a current, and a
changing magnetic field in an inductor is a voltage. Both fields are present if
there is any electri cal activity. These fields exist inside components, but they
also exist around all conductors involved in electrical activity.
In the circuits we have studied, the electric and magnetic fields outside of the
components can be ignored. But when the frequencies extend beyond 1 or 2 MH,
or if high accuracy is required, these fields cannot be ignored. When circuits
extend between remote points, then fields cannot be ignored. Trying to draw
circuits that describe the activity of fields is almost impossible. Getting a broad
understanding of how fields behave is not difficult, but it can be elusive. Some
of the problems encountered in building practical circuits involve electric fields.
For this reason, it is important to appreciate how fields and circuits relate to each
other.
Electrical Transport
All electrical activity involves both the electric and the magnetic field. The
power that is sent to us by the utility company is carried not on the power wires,
but in the fields between the wires. This may come as a surprise, but it is the
only explanation that can be used to understand the many problems that do
occur. In all circuits, fields carry all power and all signals. The conductors are
guides that direct the path of signal or energy flow It is difficult to use this
viewpoint to design circuits, yet if we do not embrace it on some level, many
interference processes will stay a mystery.
Electric fields store energy. Nature is always looking for ways to let energy
flow to a lower state. An analogy with water may be helpful here. Water in a
water tower stores energy of position, and will flow down to earth if there is a
path. If the path involves turning a turbine, then the stored energy can do work.
In the same way, stored electric field energy will follow conductors as a way to
reach a lower energy state. The energy stored in a capacitor will heat a resistor if
the resistor shunts the capacitor. A television signal is a field traveling in space.
A television antenna provides a path so that some of this energy will flow to the
receiver circuit. The field takes this path because it is flowing to a lower energy
state.
The space around us is filled with fields. These fields include utility power,
radio stations, television, radar, cell phone signals, telephones, and ham radios.
Lightning and static discharge generate fields. As these fields travel in space,
they can couple energy to conductor pairs and follow these paths into circuits.
This is nature running downhill. A circuit operates on this same principle.
Energy from the power supply is running downhill through the components to do
our bidding. There are power leads and signal cables connecting to our circuits.
Nature will couple field energy to these conductors. The coupled energy goes in
both directions on every conductor pair, and some of this energy can enter our
circuits. A power cable can carry power into a circuit, and at the same time it can
transport interference in or out of a piece of hardware. The interference field can
be transported between the power conductors or between the power conductors
and earth.
The simplest shield is a metal enclosure. The enclosure can take on many
shapes; often it is a woven braid that extends over wires in a cable. The shield
material can be aluminum, steel, or a conductive paint. If the shield surrounds a
circuit, the electric fields associated with voltages in the circuit will be confined
to the enclosure. Electric fields that are external to the shield will stay outside.
Inside the enclosure, the effect of the electric fields can be equated to
capacitances from the circuit to the shield surfaces. These capacitances function
just like circuit components, allowing current to flow. The shield is usually
connected to the circuit common to avoid unwanted coupling and capacitive
feedback. If the shield is not connected and left floating, it functions as a
capacitive divider and external fields can add signal to the circuit.
In an ideal shield enclosure, the fields outside the box are reflected and do not
enter the enclosure. Of course, the ideal perfect enclosure rarely exists. Fields
enter via the power leads and cross through the transformer coils. Fields enter on
input and output leads. At high frequencies the fields can enter through seams
and holes. Shielding that is effective at 60 Hz against the electric field may be
ineffective against a magnetic field.
In many circuits it is only necessary to shield the input signal right up to the
input of the circuit. This is electric field shielding. The rest of the circuit is often
relatively insensitive to coupling. If the input stage adds gain and has a low
output impedance, then coupling to the second stage can be ignored. It is often
desirable to add a small RC filter right at the input terminals as added protection.
Typical values might be a series RC circuit consisting of a 10042 resistor and a
100-pF capacitor.
Integrated circuits have the advantage of being very compact. This means that
the component itself is relatively immune to external fields. This places the
coupling burden on supporting components and on the cables that carry signals
in and out of a circuit.
Common-Impedance Coupling
The power supply circuits that we have considered have diode rectifiers and
electrolytic capacitors. It is not uncommon for currents in the electrolytic
capacitors to be several amperes. This current can cause a voltage drop in the
wiring. As a result, the common side of the capacitor may not be at the same
potential as the centertap of the transformer. If a conductor has a resistance of 10
mQ, a peak current of 2 A will develop a peak signal of 22 mV. If this signal is
added to an input signal and amplified, the resulting hum can be very
objectionable. Adding more capacitance will not cure the problem.
The power supply should be built so that the leads that carry the filter current
are not a part of any signal circuit. Figure 7.1 shows a proper and an improper
connection to a power supply.
In a power amplifier output stage where there is feedback, the voltage drop in
a common lead can actually change the feedback signal. In some cases this
unwanted feedback can result in an instability. This problem can occur in output
circuits that supply high current. The solution involves wiring the circuit so that
the feedback signal does not sense the load current. Figure 7.2 shows a proper
and an improper feedback connection in a power amplifier output.
Star Connections
The conductors that are considered at 0 V can be numerous. The connections can
include an input shield, an output shield, an input common, an output common,
emitters, a transformer centertap, electrolytic capacitors, feedback connections,
filter connections, a safety ground, and connections to the chassis. There are
many ways to connect these grounded conductors together. One approach is to
select a single point for grounding and bring all of these conductors to this
common point. This is called a star connection. The idea is to eliminate any
common resistance in the paths between various grounds. If there are no external
grounds except the safety connection, then this idea can work. If the input shield
is grounded at a remote point, current can flow in the input shield to the safety
ground, and the star connection does not work. As indicated earlier, shield
current can couple interference to an input signal if the shield is one of the signal
conductors.
Transmission Lines
In the circuits we have considered, all power and signals are transported between
two conductors. The signal is the potential difference between the input lead and
the signal common. Power is also the potential difference between two
conductors. As we have seen, power and signal can sometimes share the same
conductors. Every pair of conductors is a potential transmission line. The power
and signal voltages move so fast that for all practical purposes the time of
transmission can be considered zero. But there are many situations where this
viewpoint is inadequate. For this reason, a brief discussion of transmission lines
is important.
The simplest transmission line, shown in Figure 7.5, consists of two parallel
conductors and a load resistor. We are interested in what happens the moment
the battery is connected to the line. We know that after a few microseconds the
power supplied to the resistor is V z/R. Before this can happen, field energy
must travel along the transmission line to get from the battery to the resistor.
The capacitance per foot of line determines the charge that must be supplied
to the conductors in that first nanosecond. Assume the capacitance is 10 pF per
foot. If the voltage is 1 V, the charge q in the first nanosecond is 5 x 10-12 C.
This means a current must flow equal to q/t = 5 mA. In the second nanosecond
the field has moved a total of 1 foot. The same current flows in that second
nanosecond. The current that flows implies that a magnetic field is associated
with the moving electric field. This combination of electric and magnetic field is
called a wave. The wave front moves down the transmission line at about one-
half foot per nanosecond. The current supplied by the battery is steady. This
means that the battery reacts as if a resistor of V/I = 200 S2 were placed across
its terminals. This value is called the characteristic impedance of the line.
If the line is 10 feet long, the wave reaches the terminating resistor in 20 ns. If
the terminating resistor is 200 0, the battery continues to supply 5 mA into the
transmission line until the switch is opened. The energy that leaves the battery is
converted from chemical energy to field energy before it is carried and
dissipated in the resistor. It is important to realize that the fields carry the energy
from the battery to the resistor. The wires direct where this energy is to flow.
The wires do not carry the energy. The field pattern along the transmission line
is shown in Figure 7.6.
The fields associated with the transmission line in Figure 7.6 extend out into
space. Some of this field energy escapes and does not return to the circuit.
Coaxial Cable
Another kind of transmission line geometry is the coaxial cable. The two
conductors involved in transmission are the center conductor and the outer
cylinder. This geometry is shown in Figure 7.7.
In this transmission line the fields cannot leave the confines of the cable.
Open wires are acceptable where the intent is to radiate the energy. The open
parallel wires will also accept higher operating voltages than a coaxial cable.
The coaxial transmission line looks a great deal like a shielded cable. When it
is used as a transmission line, the currents associated with the movement of the
field flow on the inner surface of the sheath. This means that both ends of the
sheath must be connected to allow this current to flow. If the cable is used for
shielding against external electric fields, the outer covering may be connected
(grounded) at one end.
Energy is now traveling in both directions at the same time. Because of the
open circuit, the wave that is returned must cancel the current as it progresses
back along the line. The voltage of the reflected wave adds to the voltage of the
forward wave. An oscilloscope placed along the line would show double the
battery voltage during the period of the first reflection. The battery continues to
supply energy to the line until the reflected wave returns to the battery. At the
battery, the returning voltage is incorrect and the wave is again reflected. This
time the battery is shut off as the second reflection sends the first wave forward a
second time.
If there were no losses in the transmission line, the energy would continue to
travel back and forth forever. What really happens involves heating losses,
radiation, and wave distortion. After a few round trips the reflections attenuate
and the steady battery voltage appears along the entire line.
If the transmission line is terminated in a short circuit, the first reflected wave
cancels the voltage at the end of the line. Wave energy is still returned to the
source in the reflected wave. (Energy cannot be dissipated in an open or short
circuit.) This time, when the returned wave reaches the voltage source, the
battery must add more energy to the wave. The second reflection requires a
battery current that is three times the initial current. Upon the fourth reflection,
the current is multiplied by five. It is obvious that the current builds on each
return reflection until a fuse blows or the conductors melt. This is the way
current builds in a short circuit.
Sine Waves and the Transmission Line
There are many conductor pairs that nature can use. These conductors include
the earth, shields, metal surfaces, conduits, oceans, and any open conductors, as
well as our circuit conductors. Nature never reads labels, color codes, or
directions. The available conductor pairs are not necessarily parallel, nor are
they properly terminated. Some of them are brought directly into our hardware,
and others add to the general ambient activity. The result is that coupled wave
energy reflects and bounces around these conductors in a very complex and
uncontrolled manner. It is much like the way light reflects and is absorbed by
objects inside a room. The patterns are extremely complex.
Fortunately, most of the circuits we build cannot respond to these fields
because the fields are changing so rapidly. Circuits are built to respond only to a
particular portion of the coupled energy and ignore the rest. If we were building
a cell phone, for instance, the circuit would be designed to respond only to the
signal that carries the information the phone needs to function. This is the beauty
of filters and resonant circuits. It allows the circuit to be very selective in how it
responds.
Power lines that cross the space between buildings can couple to any radiated
energy sharing the same space. The power conductors entering a circuit carry
any and all types of signals. They can carry this signal right down the center of
the conduit. These signals cannot do any work for us, but they can add to the
interference that enters our circuits. The signals riding the power signals can be
reflected or absorbed by line filters, provided the filters are properly installed.
Field Coupling
The most important tool in avoiding field coupling is to avoid a coupling area.
Two conductors spaced 1 inch apart will couple to field energy proportional to
the spacing and to the length of the run. To reduce the coupling, simply move
the two conductors closer together. If the length of the run can be reduced, this is
in the right direction.
If the fields are intense, then the next line of defense is to place the circuit
inside a metal enclosure. A shield around a cable or around a circuit can be very
effective. If the unwanted field enters the cable at either end, the shielding will
be ineffective. Interference can use the space inside a cable for transmission just
as easily as a desired signal can. If shielding is to be effective, all the points of
entry must be considered. The problem is similar to the leaky boat. If some of
the holes are plugged, the boat will still sink. All of the holes must be
considered.
Here is a short story to illustrate the nature of the cable problem. A circuit
consists of a sensor, a cable, and an amplifier. The connecting cable is shielded,
and the shield is bonded to the metal boxes that contain the sensor and the
amplifier, yet every time a nearby relay is operated, the circuit malfunctions.
What was the design error? It turns out to be quite simple. A wire was placed
inside the cable that connected the two metal boxes together. When this wire was
removed, the circuits functioned properly.
But why did this wire cause a problem? Because the relay contact that opened
created a field that used the cable and the earth as a transmission line. The
current in this loop would normally stay on the outside of the shield used to
cover the connecting cable. By adding a wire inside the cable, some of current in
this loop used the wire. This created a field inside the cable that coupled to the
sensor signal pair. When the relay opened, the sensor line sensed a signal that
was coupled to the amplifier. By removing the wire, the interference field stayed
outside the cable.
Radiation
The half-dipole antenna is not the only conductor geometry that radiates. The
other basic geometry is a loop of wire. If the current loop is a quarter-
wavelength in diameter, the loop will radiate energy very efficiently. The high
current requirements make this a less desirable antenna design. This is another
reason why loop areas in a signal path are undesirable. They can function as
radiators.
Circuits that do not confine their fields will radiate. The radiators might not
be efficient, but they can interfere with nearby pieces of electronics. The main
reason the FCC requires radiation testing on electronic devices is to make sure
they do not interfere with radios and television reception. The standards used in
Europe go far beyond this requirement. Circuits will not radiate efficiently
unless there is frequency content above 1 MHz. This can occur with certain
types of switching regulators or with circuits that are oscillating. When there is
arcing, there is usually radiation. Digital circuits with their high-speed logic can
also radiate.
In Conclusion
This last chapter has touched upon many new areas, and includes topics that
form the subject of many different books. If your interest is caught by this kind
of discussion, you might enjoy two more of my books, The Fields of Electronics
and Grounding and Shielding, 4"' Edition (both published by John Wiley &
Sons). They are written for electrical engineers, but you should now have
enough experience with basic electronics to be able to read them if you wish.
If you've enjoyed building the Learning Circuits and found the discussions
interesting, electronics could be a good career choice. However you decide to
pursue your interest in electronics, I hope this book has given you a good start,
and that you've had some fun along the way.
Introduction
The story of electricity starts with the electrical forces that hold all atoms
together. One of these forces is the electric force. This force exists between the
electrons and protons in every atom. Molecules are formed when outershell
electrons are shared between atoms. In metals, atoms are packed tightly together
and the electrons are not shared. For this reason, the outershell electrons can
easily move between atoms. These are known as free electrons. Materials with
free electrons are called conductors. The most common conductors used in
electronics are copper, aluminum, and iron. Materials whose electrons cannot
move freely are called insulators.
Electrons have two qualities: an electrical charge and an electric field. The
electric field is a region of influence (force) that can move other electrons.
Within an atom the fields from the electrons are exactly balanced by the fields
from the positive charge of the protons in the nucleus. When there are extra
electrons, their electric fields combine to form a larger field. This field extends
outside the atoms into the space around the object.
When two insulators are rubbed together, charges can rub off one surface and
deposit on the other. For example, rubbing a silk cloth over a plastic rod will
cause some electrons to move from the cloth to the rod. Once electrons are
added to the rod, they can be transferred to other objects through a contact.
If two conductors that hang from a string are touched, they will repel each
other. This repelling force is attributed to the electric fields that surround these
added electrons. This repelling force field acts at a distance on every electron in
the vicinity. Free electrons in nearby conductors will move as a result of this
force.
The absence of electrons also has an electric field. The absence of electrons is
equivalent to a positive charge. Two objects with a positive charge will repel
each other. Objects with a negative charge will attract objects with a positive
charge.
If a negatively charged object is brought near a conductor, the free electrons
on the conductor will be influenced by the external electric field and move away
from their points of equilibrium. These electrons will move until there is a
balance of forces between the displaced electrons and the external field. The
electrons that move away from a part of the surface leave behind a field that for
all practical purposes behaves like it is associated with positive charges. These
pseudo-positive charges have the same mobility as actual electrons. It is possible
to have areas of positive and negative charge on the same conductor.
The unit of charge is the coulomb, abbreviated C. The letter q or Q is often used
to represent a given charge. One coulomb of negative charge is 6.28 x 1018
electrons. In most circuits the amount of charge involved is expressed in gC,
which is 10-6 coulombs.
The field that surrounds a charged object is called an E or electric field. This
field can exert a force only on another field. To measure the strength of a field, a
second charged object must be used. The charge on the second object must be
very small compared to the initial charge, or the field being measured will be
modified. The field is measured by noting the force on this small test object. At
every point in space the force has a direction. If arrows are used to represent the
direction of the force, the stems of the arrows can be interconnected to form
electric field lines. By convention, field lines start on positive charges and
terminate on negative charges. These field lines can be used to map the character
of the field between charged objects.
Work and Energy
Energy is stored in capacitors and inductors. This energy can do work by heating
a resistor or turning a motor. Voltage is defined in terms of work. Before we can
define voltage, we must define work.
The definition of work is "force times distance." The direction of the force
must be along the path of motion. When a mass is lifted on Earth, the work is
equal to the mass times distance. When 1 kilogram of water is lifted 1 meter, the
work done on the water is equal to 1 kilogrammeter. This work is stored as
potential energy or energy of position. Potential energy has the same units as
work. The energy stored by the water can be used to do work by releasing the
water and letting it turn a water wheel.
A force acting on a mass can also accelerate that mass. An example is a car
on rails. The work done on the car is again force times distance. In this case the
motion of the car stores energy. Energy of motion is called kinetic energy. The
units of kinetic energy are the same as the units of work. This energy of motion
could be converted to energy of position by letting the car roll up an incline.
In Figure 8.1, the E field is uniform and the force on the test object is constant.
The work required to move the test object between the two conductors is f x d
where d is the spacing. The work W is therefore
In most circuits, the E field is not a part of the calculation. However, you
should realize that charges respond to the E field and not to voltage.
Voltage (Electrical Pressure)
The definition of voltage involves an electric field and charges on the surface
of the conductors. If there is a voltage difference, then a field and a charge
distribution must be present on conductors. The two go together and cannot be
separated.
The voltages in electronics can cover a wide range, and many units are used
in order to cover this range. The millivolt or mV means 0.001 V. The microvolt
or tV is 0.000001 V. The kilovolt or kV is 1,000 V. The megavolt or MV is
1,000,000 V.
Current
Electrons are attracted to a positive potential. It would seem natural to say that
the direction of current flow is the same as the direction of electron flow.
Historically, however, the opposite convention was established, and it has
persisted. In all circuit analysis, current is said to flow from a point of positive
potential to a point of more negative potential. Electrons flow in the opposite
direction, to points of higher positive potential.
A resistor is a component that limits the flow of free electrons. Resistors play a
central role in all electronics. When a voltage is placed across the terminals of a
resistor, the electric field in the resistor causes electrons to move. The average
electron velocity will depend on the materials used to form the resistor.
The relationship between voltage, current, and resistance is Ohm's law. This
relationship can be written in three different equations:
The common carbon resistor covers the range from 10 S2 to 22 MU, which is
22,000,000 Q. The Greek capital letter omega (S2) is the abbreviation for ohm.
The other standard abbreviations are kQ for 1,000 ohms, MS for megohm, and
mS2 for thousandths of an ohm.
where Pis power in watts, W is work in joules, and t is time in seconds. The
work in moving a charge through a potential difference was given by W = qV. If
both sides of this equation are divided by time, then W/t equals power P and q/t
equals current I. This means that
The power dissipated in a resistor can be determined from Equation 8.9 if you
substitute current or voltage from Ohm's law. Since V = IR or I=V/R
Capacitors
The use of the letter C for both capacitance and charge can be confusing.
Which meaning is intended can always be inferred from context.
The circuit symbol for a capacitor is:
The voltage across the plates of a capacitor implies that there are charges on
the plates and that there is an electric field between the plates. If the charge
increases linearly, then the voltage also increases linearly. If both sides of
Equation 8.11 are divided by time t, the term Q/t is current and the equation
becomes
This equation states that a steady current flowing into a capacitor results in an
increasing voltage. If 1 A flows into a 1 F capacitor, the voltage rises at 1 V per
second. If 1 µA flows into a 1 tF capacitor, the voltage rises at 1 V per second.
When V/t changes, the current adjusts to the new value.
The conductors shown in Figure 8.1 form the plates of a capacitor. To calculate
the energy stored in this capacitor for a given voltage, we divide the stored
charge Q into many smaller charges q. The plan is to move the charge across the
space in small increments. When the charge on the plates is 0, it takes no work to
move the first charge from the top plate to the bottom plate. The second charge q
takes a small amount of work, as there is now an electric field. From Equation
8.3, the work required to move the last charge q is qV. The total work is the
average work, or
This equation can be related to the capacitance by substituting Q from Equation
8.11:
Capacitors store energy in the electric field between the plates. This energy is
not dissipated. It is stored as potential energy-that is, energy that can do work
later. The ideal capacitor will hold stored energy until a path is provided for the
charge to leave. A parallel resistor can provide this path. The stored energy will
simply heat the resistor.
Practical Capacitors
The plates in Figure 8.1 would form a very small capacitor. Capacitance
increases with larger surface area and reduced spacing between the plates. The
capacitance also increases if the space between the plates is filled with an
insulating dielectric rather than air. Typical dielectrics are glass, mica, Mylar,
and polypropylene. The increase in capacitance brought about by using one of
these materials is called the dielectric constant.
A magnetic field can be described two ways. One way is the induction or B
flux, and the second way is the H flux that is created by current flow. The
changing induction flux induces the voltage stated in Lenz's law. For a given
current, the induction flux can be increased by placing magnetic material in the
flux path. This is exactly analogous to the increase in charge storage when a
dielectric is added to a capacitor. The ability of a material to increase the
induction or B flux is called permeability. Iron has a permeability that can
exceed 10,000. This property of iron makes transformers and motors practical at
60 Hz. By having iron in the magnetic path, the current required to establish the
induction flux can be held to practical limits.
This energy is stored in the magnetic field. The energy is constant as long as the
current is sustained. In practice, however, a sustained current can only occur in a
superconductor where the resistance of the coil is 0. If the energy stored in a
capacitor is considered potential energy, then the energy stored in an inductor by
a current can be considered kinetic energy. The moving charges store the energy.
Transformers
Semiconductor Materials
If boron is used as the dopant, then the material becomes p-type silicon.
Boron has one less electron than silicon, so instead of providing a free electron,
it provides a receptor site. The absence of an electron is called a hole. P-type
silicon is also a conductor, as the holes behave very much like the pseudo-
positive charges on a conductor.
A sandwich of pnp or npn material forms a transistor. The outer two materials
are called the emitter and the collector. The center material is called the base. If
an electric field is impressed across the sandwich, the two diodes inhibit current
flow in either direction.
In an npn transistor, the electrons flow in the forward direction through the
base emitter diode, and then the sandwich acts like all n material. Electrons flow
across from the emitter to the collector. This assumes there is a collector voltage.
There is a multiplication effect by which a small amount of base current results
in a significant amount of emitter to collector current. This multiplication of
current is called transistor action. In a pnp transistor the exact same
multiplication occurs, except that all the current directions are reversed. This
action of transistors makes them very important components in electronic
circuits.
The angles that are frequently used in electricity are 45°, 90°, 180°, and 360°.
When measured in radians, these are n/4, n/2, it, and 2n radians.
Frequency
The unit of frequency used in electronics is the hertz. One hertz is equal to one
cycle per second. The cycles can be of any recurring event. The abbreviation for
hertz is Hz. The expression kHz is read kilohertz and means 1,000 cycles per
second. The abbreviation MHz is read megahertz and the expression GHz is read
gigahertz. One GHz is equal to 1,000 MHz.
Sine Waves
Sine wave voltages and currents play an important role in all electronics. Circuit
analysis uses sine waves, as this is the only waveform that remains unchanged
throughout most circuits. Sine waves are also used as carrier signals in most
communications channels.
The sine wave and a rotating radius are shown in Figure 8.2.
Figure 8.2 The sine wave function using a rotating radius
If the angle in Figure 8.2 increases with time, the value of h will move
between + and -1 during every revolution. If the angle increases 360 degrees per
second, then the height at any time t is equal to
If the angle makes two revolutions per second, then the frequency f is 2 Hz The
height h for any frequency f and for any value of time is
We can use this sine function to represent any parameter that changes
sinusoidally. For example, a sinusoidal voltage v is represented by the
expression
The voltage v reaches a peak value of + V once per cycle. If f = 20 Hz, then v
varies between + V and -V 20 times per second.
where V is the peak value of voltage, f is the frequency in Hz, and t is time in
seconds.
In circuits with resistors, capacitors, and inductors, the voltages and currents
peak at different times. This is represented by pointers that are separated by
phase angles.
When a sinusoidal voltage is placed across a capacitor, the current that flows
depends on how rapidly the voltage is changing. For a sine wave voltage, the
voltage changes most rapidly at the zero crossing of voltage. The maximum rate
of change is 2nf Vp where f is frequency and Vp is the peak voltage of the sine
wave. The peak current that flows from Equation 8.12 is
The current that flows is also a sine wave where Ip is the peak value. The ratio of
peak voltage to peak current is called the reactance of the capacitor and is
The current in a capacitor for a sine wave voltage is always maximum when
the voltage is 0. Figure 8.3 shows the timing relationship between current and
voltage in a capacitor. Using the pointer system, you can see that the current
leads the voltage by 90 electrical degrees. This relationship holds for all
capacitors at all frequencies. Energy is stored in a capacitor when there is
voltage. This means that peak energy is stored twice per cycle. Energy is stored
in a capacitor; it is never converted to heat in a capacitor.
Figure 8.4 uses the pointer system to show the timing relationship between
voltage and current in an inductor. If the voltage is selected as the reference
pointer, then the current pointer points straight down. This means that the current
in an inductor lags the voltage by 90 electrical degrees. This relationship holds
for all inductors at all frequencies. Energy is stored in an inductor twice per
cycle when the current is at maximum. Energy is stored in an inductor; it is
never lost as heat.
Figure 8.4 The timing relationship between current and
voltage in an inductor
The abbreviation dc stands for direct current. Direct current means that the
voltage or current does not vary. A battery voltage is an example of dc voltage.
If one ampere is drawn from the battery, this current is also called dc. Although
the word "current" is present in the abbreviation, it is not always appropriate, as
dc refers to both voltage and current.
The abbreviation ac stands for alternating current. The power voltages that we
use are 60 Hz ac. Alternating current is a varying voltage or current. Again, the
word "current" in the abbreviation may not be appropriate.
Voltages and currents are usually referred to in terms of their heating ability.
At dc, the power dissipated in a resistor is V2(dc)/R. It is very convenient if an
ac source with the same measure provides the same heat. In other words,
V2(ac)/R should provide the same heat. But when an ac voltage heats a resistor,
the power varies during the cycle. The average squared value of a voltage is one-
half the peak value. Thus we could say that power is equal to Vp2/2R. Rather
than involve this factor of two, the accepted practice is to refer to sine waves in
terms of their peak value divided by the V. This is their rms, or root mean square
value. In this book all the ac voltages and currents are assumed to be rms values
unless otherwise stated.
If V is the rms value of a voltage, the peak voltage is 1.414 V. The peak-to-
peak value is 2.818 V. In our homes the rms value of voltage is 120 V. This is
169.7 V peak or 339 V peak-to-peak.
Square Waves
Useful square waves must transition between voltages in a very short time.
These transitions are called rise and fall times.
The rms value of a symmetric square wave is the peak value. If the square
wave is 0 half the time, the heating value is one-half. This makes the rms voltage
the peak value divided by the square root of two.
Square waves are a valuable tool for testing circuits. A square wave is made
up of many sine waves. By using a square wave to test a circuit, the response can
yield information not available by using a signal at one frequency.
Decibels
In the early days of the telephone, a unit was invented that represented the
smallest detectable change in sound power level. This unit was called a decibel.
The term bel was used to honor Alexander Graham Bell, the inventor of the
telephone. The bel is the logarithm of the power ratio. The decibel was defined
as 10 log P1/P2 where P, and P2 are these power levels. When measured as
voltages, this power ratio becomes 20 log V,/ V2. For 1 decibel, the ratio of
voltages is equal to 1.122. This assumes the voltages are applied to the same
resistance. This logarithmic measure of ratios has come into common use, and
the original power definition is sometimes lost. A circuit with a gain of 100 is
said to have a gain of 40 decibels, abbreviated 40 dB. This measure has nothing
to do with power.
The beauty of the decibel in describing performance is that decibels add while
gain factors multiply. For example, if two circuits have a gain of 23 dB and 12
dB, the combined gain is 35 dB. If a circuit attenuates a signal by 45 dB and the
gain following is 60 dB, the overall signal gain is 15 dB. If the initial signal was
5 dB volts, the resulting signal would be 20 dB volts or 10 V The dB makes
calculations very simple. See the table showing dB values.
Frequency Response
The components you will need in the Learning Circuits are available in most
electronics parts stores. Some stores carry loose stock, as opposed to
individually packaged components. The advantage to loose stock is lower cost.
Individually packaged items will be more clearly identified as to value and
rating. As an example of cost, X W carbon resistors in bulk cost a manufacturer
under two cents each. A single packaged resistor may cost 50 cents or more.
This cost can add up if you need a hundred resistors.
The number of components you will need is not exact. You may want to leave
some of the circuits assembled rather than tear them down for the next lesson, or
you may want to experiment with constructions that are not a part of the
Learning Circuits. This is entirely up to you. You may want to start slowly and
buy a few parts. Buying the parts is part of the fun. It is also part of learning
about electronics.
You will probably purchase resistors that are ! W 20% carbon, although metal
film resistors can be used. In many cases the exact values indicated are not
critical and a nearby value will be acceptable. Except for size, 1-W resistors will
work fine.
The smaller capacitors can be metalized Mylar, although other types will be
quite adequate. The larger capacitors (above 10 .tF) should be electrolytics rated
35 V or higher. In the voltage doubler circuit, the capacitor must be properly
rated.
The power diodes 1N1002 can be any diode in the series 1N1001 through
1N1005. The last number defines the reverse voltage, and in our circuits this
voltage is quite low. There are many other power diodes that would function in
these circuits.
The transistor types suggested are readily available, although many other
types would be acceptable. The best way to substitute another type is to look at
specification sheets. There is often help available in most stores as to how you
can make substitutions. Manufacturers provide component data on the Internet.
You may want to try this approach. Just type in the part number on "Google" and
you will be surprised. The only things that are important are the maximum
voltage and current ratings and the gain, called P.
There are many ways to build experimental circuit boards. I have suggested
an epoxy board with a grid of punched holes. Packages of pins are available that
press-fit into these holes. Components can then be soldered to the pins. You
might want to buy a package of 100 pins to get started. A tool to press-fit the
pins into the board is often a help. The following list of components will allow
you to build all the circuits in this book.
6 10 W, 2 18 kS2, 2 22 k12
Potentiometer
Capacitors:
The majority of these components can be metalized Mylar rated greater
than 50 V. Other dielectrics can be used. The voltage ratings should be 50 V
or greater.
1 100 pF, 2 0.001 .tF, 4 0.01 .tF, 3 1.0 p.F, 4 100 µF 35-V electrolytic
Semiconductors
1 LF353 IC amplifier
Inductor
Hardware
1 SPDT switch
test board with a grid of holes approximately 6" by 8". An epoxy board is
recommended.
1 package of 100 press-fit pins (must match the hole diameter in the board)
1 18-V ac adapter rated X ampere or greater; #PHC-AC-1888C or equal
needle-nose pliers
wire cutter
Soldering
You will need to know how to solder if you want to construct the Learning
Circuits in this book. Soldering is the process of heating an alloy of tin and lead
called solder and using it to make an electrical connection. It requires a supply of
solder and a soldering iron to provide heat.
There are many types of soldering irons. The irons used for electronics are not
the 100-or 200-W variety. These heavy irons can damage sensitive components
through overheating. In electronics we use a smaller iron that has a small pointed
tip, ideally one with a regulated temperature at the tip. 20 or 30 watts is
adequate.
The solder you will use is usually provided on a spool. It looks like a coil of
white uninsulated wire. In the center of the solder strand is a material called
rosin. During the soldering process the rosin aids in cleaning the surfaces to
allow the solder to flow.
After your soldering iron is hot, make sure the tip can melt solder. If you have a
new iron, melt solder on the entire tip area. Wipe off any excess solder with a
heavy cloth. If you leave a soldering iron on for a period of time, a layer of
oxidation or scum will form. Always wipe this layer of scum off the iron with a
rag before you begin soldering. Heat cannot flow effectively through this scum.
The only way heat can properly flow from the soldering iron is through the
melted solder on the tip. The tip is ready when it has a thin, bright white layer of
melted solder on its surface. The heat from the soldering iron must flow to the
conductors being soldered. Touch the soldering iron tip to the conductors to be
soldered. Heat will flow to the conductors through the "tinned" tip. Feed the new
solder by hand to the leads being soldered near the point of contact, not to the
soldering iron. When the leads are hot, the solder will melt onto the leads.
A good solder connection is a thin layer of solder that connects two or more
conductors. It is not just a glob of solder. The solder should flow so that the
solder tapers to the lead. Any foldback of solder is suspicious. These poor
junctions are called rosin joints. They can be a source of trouble. A rosin joint
can be intermittent or even an open connection. This is a problem you do not
need when you are getting started.
Soldering Hints
If you have done soldering in other applications, such as plumbing, you may
have used acid flux to help with bonding. Acid flux is never used in electronic
soldering. The acid can eventually do damage to components. The solder used in
electronics contains a rosin that forms the core or center of the solder The rosin
helps clean the surface and is adequate to do most circuit soldering.
One last tip: Avoid using a solder joint as a mechanical support, except on a
very temporary basis. The right way is to use a mounting pin or terminal so that
the connection is anchored to a circuit board. Solder is an electrical connection,
not a mechanical support.
However you obtain this equipment, you will learn a great deal from using it.
There is truly no substitute for hands-on experience in electronics, and this is
what I have tried to give you with the Learning Circuits. Have fun with them,
and your understanding of electricity will grow very quickly.
Introduction
To read the text of this book and to work the problems, you need a basic
understanding of the branch of mathematics called algebra. The tools of algebra
enable us to do all the arithmetic operations without requiring us to use specific
numbers. This generalization gives algebra tremendous flexibility and
usefulness.
I have used algebra only where I felt it was really necessary, and this
discussion is not meant to be a comprehensive explanation of algebra, but only a
review of those topics I actually use.
If you have already studied algebra, this brief section will refresh your
memory. If you have not studied algebra before, this section should give you a
sufficient understanding to read the text. The only way to gain proficiency in
using algebra is to work many problems. Unfortunately, there is not space in this
book for those problems. If you want to really learn algebra, which I
recommend, there are many textbooks available. A book in this series called
Practical Algebra is recommended.
The operations used in algebra are essentially the same ones used in arithmetic.
The common operations are addition, subtraction, multiplication, and division.
The symbol for addition is the + sign and the symbol for subtraction is the - sign.
Simply placing two quantities next to each other indicates that they are to be
multiplied. Sometimes the symbol x is used to indicate multiplication. In
algebra, division is usually indicated by a horizontal line. (The division symbol -
used in arithmetic is not very convenient.) Items below the line are divided into
items above the line.
The difference between arithmetic and algebra is that algebra uses symbols in
addition to numbers. In arithmetic, you might write 2/3, which means "two
divided by three." In algebra you can perform the same operation and write it the
same way, but use the symbols a and b instead of the numbers. The term alb
means "a divided by b."
The symbol = means "equals." In this example the letter v stands for velocity
and the letter t stands for time. For example, if you drive 60 miles per hour for 2
hours the distance is 120 miles. In this case v = 60 miles per hour and t = 2
hours. In words, 120 miles equals 60 miles per hour times 2 hours. The
statement that v = d x t or v = dt is very compact and covers all velocities,
distances, and periods of time for any system of units. Distance, velocity, and
time are all called parameters in this statement.
The simplest representation of the numbers we use places them on a straight line.
The center point is 0. To the right are the increasing positive numbers and to the
left are the increasing negative numbers. The addition of numbers simply adds
lengths of the line from the origin. The length 3 is added to the length 5 for a
value of 8. If both numbers are negative, the length is given a negative direction.
The subtraction of numbers is the distance between points on the line. The
distance between 6 and 8 is plus 2. This is written as 8 - 6 = 2. The distance from
8 to 6 is a minus 2. This is written as 6 - 8 = -2. The multiplication or division by
a negative number reverses the direction of the original number. For example,
minus 1 times a plus 6 is a minus 6. Minus 6 times minus 1 is a plus 6. A minus
6 divided by a minus 1 is also a plus 6. When symbols are used instead of
numbers, they are all treated as if they were positive values. For example, (-2) (-
a) equals +2a. If a turns out to be a negative number, the answer will be
negative.
About Parameters
Algebra has some further rules that must be followed, or errors can result.
The proper sequence of operations must be followed. The statement "three plus
five times two" can be read two ways. The first way is to add three and five
together and then multiply by two. The second way is to multiply five by two
and then add three. The first answer is 16 and the second answer is 13. Algebra
uses parentheses to indicate sequence.
If the first answer was intended, then the three and five must be placed in
parentheses. The rules of algebra require that terms inside the parentheses must
be treated as one value. In symbolic form, (a + b)c is different from a + bc. Note
that (a + b)c can also be written as ac + bc. Of course, all of these rules can be
checked using numbers.
Solving an Equation
Consider the equation d = vt. We can divide both sides of the equation by t and
obtain the result d/t = v. We can do this by drawing a line under both sides of the
equation. We then place tin the denominator on both sides of the equation. We
do not know the value of d/t, so we leave it in symbolic form. On the other side
of the equation, vt/t is the same as v because t/t is the same as 1. We say that the
is cancel. We can develop a third relationship by dividing both sides by v and
obtain the equation d/v = t. These new equations are valid if the first equation is
true. We could add or subtract any parameter from both sides of the equation
without affecting the balance of values.
To see how the rules work, we can solve for R in the equation (3R - S)/ W =
C. First multiply both sides of the equation by W. The result is 3R - S = CW
Next add S to both sides of the equation. The result is 3R = CW + S. Dividing
both sides by 3, we obtain R = (CW+ S)/3. If the parentheses are removed from
the original equation, it reads 3R - S1 W = C. We can add S1 W to both sides
and obtain 3R = C + S/W. Now R = C/3 + S/3 W, a very different expression.
Algebraic Identities
The following identities are always true and perfectly general. For example, the
identity a/a = 1 can be extended to cover b/b = 1 or (c + 1)/(c + 1) = 1.
Exponents
A notation that finds frequent use in algebra is the exponent. The simplest
example of an exponent is the square. The expression a2 is read "a squared" (or
sometimes "a to the second power"), and it means a x a. Similarly, a3 ("a
cubed") means a x a x a. Exponents are additive when the same base parameter
is involved in multiplication. For example, a2 x a3 = a5. The expression a2b3
does not equal (ab)5. It follows that in division, exponents subtract. As an
example, a5/a2 = a3. A term to the exponent zero is one for all parameters. This
is because a"/a" = 1 for any value of a or n, and (n - n) is 0. The exponent 1 is
redundant, as it leaves the value unchanged; a' = a.
The term a2/a5 can be written two ways. When the exponents are subtracted,
the result can be a-3 or 1/a3. A simple rule follows: Any term can be moved
from the numerator to the denominator by changing the sign of its exponent. The
same rule applies when a term is moved from the denominator to the numerator.
As an example, the term alb-3/c 3d could also be written as cad-'/a 2b3.
Fractional exponents can be used. The simplest application is the square root.
The notation a` means the square root of a. This can also be written as N/a. The
definition of the square root is a number that multiplied by itself yields the
number. The product a'12a'/2 = a' = a.
When an exponent is used with the number 10, the 10 represents the number
of zeros following the one. For example, 102 equals a 1 followed by two zeros,
or 100;106 is 1,000,000;10-6 is one-millionth, or 0.000001. In this case the
exponent is the number of decimal places before the decimal point. The process
of adding zeroes is referred to as powers of 10. Powers of 10 are much easier to
use than large groups of zeros.
To add the terms a/b to c/d we can do the same thing. The common
denominator is bd. To change alb so that the denominator is bd, we multiply the
numerator and denominator by d. The result is ad/bd. Similarly, the term c/d can
be changed to cb/bd. With a common denominator, the sum can be written (ad +
cd)/bd = (a + c)/b.