Johnson John Johnny Johnson: Hilbum
Johnson John Johnny Johnson: Hilbum
Johnson
John L. Hilbum
Johnny R. Johnson
V
KVL
v—it
L-u *
h
c
7-
,v
. ) - lo
Qv,
n
dJl.
k /if
http://archive.org/details/basicelectriccirOOjoh_oOe
BASIC
ELECTRIC CIRCUIT
ANALYSIS
BASIC
ELECTRIC CIRCUIT
ANALYSIS
Johnson, David E
Basic electric circuit analysis.
Includes index.
1. Electric circuits. I. Hilburn, John L., 1938-
10 9 8 7 6
PREFACE x/77
1 INTRODUCTION 7
2 RESISTIVE CIRCUITS 75
3 DEPENDENT SOURCES 49
3.1 Definitions 49
3.2 Circuits with Dependent Sources 50
3.3 Operational Amplifiers 52
3.4 Amplifier Circuits 54
vii
t/iii Contents
4 ANALYSIS METHODS 61
5 NETWORK THEOREMS 89
Contents jx
16 TRANSFORMERS 416
INDEX 575
PREFACE
This book was written for a one-year course in linear circuit analysis in the sophomore
year. Such a course is basic in electrical engineering and is usually the first encounter
of the student with his or her chosen field of specialization. It is imperative, therefore,
for the textbook used to cover thoroughly the fundamentals of the subject and at the
same time be as easy to understand as it is possible to make it. These have been our
objectives throughout the writing of the book.
Most students, when they take this subject, will have studied electricity and mag-
netism in a physics course. This background is helpful, of course, but is not a prereq-
uisite for reading the book. The material presented here may be easily understood
by a student who has had a basic course in differential and integral calculus. The differ-
ential equations theory required in circuit analysis is fully developed in the book and
integrated with the appropriate circuit theory topics. Even determinants, Gaussian
elimination, and complex number theory are presented in appendices.
The operational amplifier is introduced immediately after the discussion of the
resistor, and appears, as a matter of course, along with resistors, capacitors, and
inductors, as a basic element throughout the book. Likewise, dependent sources and
their construction using operational amplifiers are discussed early and are encountered
routinely in almost every chapter.
To aid the reader in understanding the textual material, examples are liberally
supplied and numerous exercises, with answers, are given at the end of virtually every
section. Problems, some more difficult and some less difficult than the exercises, are
also given at the end of every chapter. A been made to include a
special effort has
number of problems and exercises with realistic element values. Of course, network
scaling, which is also presented, can be used to make almost all of the remaining
problems practical. In particular, in the chapter on amplitude and phase responses,
problems are given that relate to electric filters, which, of course, are very useful
xiii
xiv Preface
circuits. Active filters, using operational amplifiers, as well as passive filters, are used
as examples. Finally, a select few of the exercises and problems are used to extend the
theory discussed in the chapters. In this way, optional material is included without
INTRODUCTION
Electric circuit analysis, in nearly every electrical engineering curriculum, is the first
course taken in the major area by an electrical engineering student. Virtually all
More complicated circuit elements may have more than two terminals. Transistors
and operational amplifiers are common examples. Also a number of simple elements
may be combined by interconnecting their terminals to form a single package having
any number of accessible terminals. We shall consider some multiterminal elements
later, but our main concern will be simple two-terminal devices.
Introduction Chap. 1
An example of an electric circuit with six elements is shown in Fig. 1.2. Some
authors distinguish a circuit from a network by requiring a circuit to contain at least
one closed path such as path abca. We shall use the terms interchangeably, but we may
note that without at least one closed path the circuit is of little or no practical
interest.
electrical element
system of units so that when a quantity is described by measuring it we can all agree
on what the measurement means. Fortunately, there is such a standard system of units
that is used today by virtually all the professional engineering societies and the
authors of most modern engineering textbooks. This system, which we shall use
throughout the book, is the International System of Units (abbreviated SI), adopted in
I960 by the General Conference on Weights and Measures.
There are six basic units in the SI, and all other units are derived from them. Four
of the basic units, the meter, kilogram, second, and coulomb, are important to circuit
theorists, and we shall consider them in some detail. The remaining two basic units are
the degree Kelvin and the candela, which are important to such people as the electron
device physicist and the illumination engineer.
The SI units are very precisely defined in terms of permanent and reproducible
quantities. However, the definitions are highly esoteric and in some cases are com-
prehensible only to atomic scientists. 1
Therefore we shall be content to name the basic
units and relate them to the very familiar British System of Units, which includes
inches, feet, pounds, etc.
The basic unit of length in the SI is the meter, abbreviated m, which is related to the
British system by the fact that 1 inch is 0.0254 m. The basic unit of mass is the kilogram
'Complete definitions of the basic units may be found in a number of sources, such as, for example,
"IEEE Recommended Practice for Units in Published Scientific and Technical Work," by C. H. Page
et al. {IEEE Spectrum, vol. 3, no. 3, pp. 169-173, March 1966).
Chap. 1 Introduction 3
(kg),and the basic unit of time is the second (s). In terms of the British units, pound- 1
mass is exactly 0.45359237 kg. and the second is the same in both systems.
The fourth unit in the SI is the coulomb (C). which is the basic unit used to measure
electric charge. We shall defer the definition of this unit until the next section when we
consider charge and current. The name coulomb was chosen to honor the French
scientist, inventor, and army engineer Charles Augustin de Coulomb (1736-1806), who
was an early pioneer in the fields of friction, electricity, and magnetism.
We might note at this point that all SI units named for famous people have
abbreviations that are capitalized. Otherwise, lowercase abbreviations are most often
used. It is also worth mentioning that we could choose units other than the ones we
have selected to form the basic units. For example, instead of the coulomb we could
take the ampere (A), the unit of electric current to be considered later. In this case the
coulomb could then be obtained as a derived unit.
There are three derived units in addition to the ampere that we shall find useful in
circuit theory. They are the units used to measure force, work or energy, and power.
The fundamental unit of force is the newton (N). which is the force required to accel-
erate a 1-kg mass by 1 meter per second per second (1 m/s 2 ). Thus 1 N = 1 kg-m/s 2 .
The newton is named, of course, for the great English scientist, astronomer, and
mathematician Sir Isaac Newton (1642-1727). Newton's accomplishments are too
numerous to be listed in a mere chapter.
The fundamental unit of work or energy is the joule (J), named for the British
physicist James P. Joule (1818-1889). who shared in the discovery of the law of
conservation of energy and helped establish the idea that heat is a form of energy. A
joule is the work done by a constant 1-N force applied through a 1-m distance. Thus
1 J = 1 N-m.
The last derived unit we shall consider is the watt (W). which is the fundamental
unit of power, the rate at which work is done or energy is expended. The watt is
defined to be J/s and is named in honor of James Watt (1736-1819), the Scottish
1
10 9 Giga G
10 6 Mega M
10 3 Kilo k
10-3 Milli m
irr 6 Micro M
10- 9 Nano n
10-12 Pico P
Introduction Chap. 1
As an example, at one time a second was thought to be a short time, and fractions
such as 0.1 or 0.01 of a second were unimaginably short. Nowadays in some applica-
tions, such as digital computers, the second is an impracticably large unit. As a
result, times such as 1 nanosecond (1 ns or 10" 9 s) are in common use. Another com-
mon example is 1 gram (g) = 10" 3 kg.
EXERCISES
1.1.1 Find the number of millimeters in 10 km. Am. 10 7
1.1.2 If a mile equals 5280 ft, how many miles in 10 km? Ans. 6.2137
1.1.3 Find the work in millijoules done by a constant force of 25 /zN applied to a
mass of 4 g for a distance of 10 m. Ans. 0.25
We are all familiar with gravitational forces of attraction between bodies, which are
responsible for holding us on the earth and which cause an apple dislodged from a tree
to fall ground rather than to soar upward into the sky. There are bodies, how-
to the
ever, that attract each other by forces far out of proportion to their masses. Also, such
forces are observed to be repulsive as well as attractive and are clearly not gravita-
tional forces.
We explain these forces by saying that they are electrical in nature and caused by
the presence of electrical charges. We explain the existence of forces of both attraction
and repulsion by postulating that there are two kinds of charges, positive and negative,
and that unlike charges attract and like charges repel.
As we know, according to modern theory, matter is made up of atoms, which are
composed of a number of fundamental particles. The most important of these
particles are protons (positive charges) and neutrons (neutral, with no charge) found
in the nucleus of the atom and electrons (negative charges) moving in orbit about the
nucleus. Normally the atom is electrically neutral, the negative charge of the electrons
balancing the positive charge of the protons. Particles may become positively charged
by losing electrons to other particles and become negatively charged by gaining
electrons from other particles.
As an example, we may produce a negative charge on a balloon by rubbing it
against our hair. The balloon will then stick to a wall or the ceiling, which are
uncharged. Relative to the negatively charged balloon the neutral wall and ceiling are
oppositely charged.
We now define the coulomb (C), discussed in the previous section, by stating that
10" 19
the charge of an electron is a negative one of 1.6021 x coulombs. Putting it
another way, a coulomb is the charge of about 6.24 x 10 18 electrons. These are, of
£Ou\o^? * \^( *
Chap. J Introduction
course, mind-boggling numbers, but their sizes enable us to use more manageable
numbers, such as 2 C, in the circuit theory to follow.
The symbol for charge will be taken as Q or q, the capital letter usually denoting
constant charges such as Q=
4 C, and the lowercase letter indicating a time-varying
charge. In the latter case we may emphasize the time dependency by writing q(t). This
practice involving capital and lowercase letters will be carried over to the other
electrical quantities as well.
The primary purpose of an electric circuit is to move or transfer charges along
specified paths. This motion of charges constitutes an electric current, denoted by the
letters i or /, taken from the French word "intensite." Formally, current is the time
(L1)
/ My>= ifrnf* '=J
The basic unit of current ampere (A), named for Andre Marie Ampere (1 775—
is the
1836). a French mathematician and physicist who formulated laws of electromagnetics
in the 1820s. An ampere is 1 coulomb per second.
3 C/s pass some specific point in the wire. This is symbolized by the arrow labeled 3 A,
whose direction indicates that the motion is from left to right. This situation is
(a) (b)
FIGURE 1.3 Two representations of the same FIG U R E 1 .4 Current flowing in a gen-
current eral element
Introduction Chap. 1
charge leaving (or, equivalently, an equal negative charge entering). Thus the current
shown entering the left terminal in Fig. 1.4 must leave the right terminal.
There are several types of current in common use, some of which are shown in
Fig. 1.5. A constant current, as shown in Fig. 1.5(a), will be termed a direct current, or
dc. An alternating current, or ac, is a sinusoidal current, such as that of Fig. 1.5(b).
Figures 1.5(c) and (d) illustrate, respectively, an exponential current and a sawtooth
current.
DC
(a)
(c)
FIGURE 1.5 (a) dc; (b) ac; (c) Exponential current; (d) Sawtooth
current
There are many commercial uses for dc, such as in flashlights and in power sup-
plies for electronic circuits, and, of course, ac is the common household current found
all over the world. Exponential currents appear quite often (whether we want them or
not!) when a switch is actuated to close a path in an energized circuit. Sawtooth waves
are useful in equipment, such as oscilloscopes, used for displaying electrical charac-
teristics on a screen.
EXERCISES
1.2.1 Find the charge in picocoulombs represented by 10,000 electrons.
Ans. 0.0016021
f(0 = 30/ :
-4/ A
Ans. 244 C
say thatfl V is 1 J/C. Jhus the volt is a derived SI unit, expressible in terms of other
units. J—
We shall represent a voltage by v or Kand use the +, — polarity convention shown
in Fig. 1.6. That is. terminal A is v volts positive with respect to terminal B. Putting it
putting it another way, energy is being supplied. To know whether energy is being
5V -5V
A O- -O B
+ v - (a) (b)
supplied to the element or by the element to the rest of the circuit, we must know not
only the polarity of the voltage across the element but also the direction of the current
through the element.If a positive current enters the positive terminal, then an external
force must be driving the current and is thus supplying or delivering energy to the
element. The element is absorbing energy in this case. If, on the other hand, a positive
current leaves the positive terminal (enters the negative terminal), then the element is
2A 2A 2A 2A
+ o- + 0-
5V 5V 5V 5V
7 7
(b) (c) (d)
Let us consider now the rate at which energy is being delivered to or by a circuit
element. If the voltage across the element is v and a small charge Aq is moved through
?, . ( <
the element from the positive to the negative terminal, then the energy absorbed
by the element, say Aw, is given by
Aw = v Aq
If the time involved is At, then the rate at which the work is being done, or the energy
w is being expended, is given by
lim -r—
Aw = ,.„
lim v
.. Aq
-j^-
Af-0A? Ar^O A/
or
dw dq
(1.3)
dt dt v
We might observe that (1.4) is dimensionally correct since the units of vi are (J/C)(C/s)
or J/s, which is watts (W), defined earlier.
i are generally functions of time, p given by (1.4) is a time-varying
Since v and
quantity. sometimes called the instantaneous power because its value is the power
It is
Summarizing, the typical element of Fig. 1.9 is absorbing power, given by p = vi.
If either the polarity of v or i (but not both) is reversed, then the element is delivering
power, p vi, to the external circuit. Of course, to say that an element delivers a
negative power, say — 10 W, is equivalent to saying that it absorbs a positive power, in
this case +10 W.
For example, the energy delivered to the element of Fig. 1.8(a) between / = and
/= 2 s is given by
2
w(2) - vr(0) = f (5)(2) dt = 20 J
Jo
Since the left member of (1.5) represents the energy delivered to the element
between t and t, we may interpret w(t) as the energy delivered to the element between
the beginning of time and t and w(t ) as the energy between the beginning of time and
t . In the beginning of time, which let us say is / = -co, the energy delivered to the
element was zero; that is,
U'(-oo) =
If t = — <x> in (1.5), then we shall have the energy delivered to the element from the
beginning up to t, given by
t
w(t) = j
vi dt
= vi dt + vi dt
J-OO J to
1 Introduction Chap. 1
which is (1.5).
EXERCISES
1.3.1 The element shown is
is absorbing
a
^
power of p = 20 W. Find the current entering
terminal b. Ans. —4 A
5V
bo-
EXERCISE 1.3.1
1.3.2 Find the energy delivered to the element of Ex. 1.3.1 between and 2 s.
Ans. 40 J
We may classify circuit elements intotwo broad categories, passive elements and
active elements, by considering the energy delivered to or by them.
A circuit element is said to be passive if the total energy delivered to it from the
rest of the circuit is always nonnegative. That is, referring to (1.6), for all t we have
The polarities of v and i are as shown in Fig. 1.9. As we shall see later, examples of
passive elements are resistors, capacitors, and inductors.
An active element is one that is not passive, of course. That is, (1.7) does not hold
for all time. Examples of active elements are generators, batteries, and electronic
devices that require power supplies.
We are not ready at this stage to begin a formal discussion of the various passive
elements. This will be done in later chapters. In this section we shall give a brief
discussion of two very important active elements, the independent voltage source and
the independent current source.
.
Chap. 7 Introduction 11
completely independent of the current through the element. The symbol for a voltage
source having v volts across its terminals is shown in Fig. 1.10. The polarity is as
shown, indicating that terminal a is v volts above terminal b. Thus if v > 0, then
terminal a is at a higher potential than terminal b. The opposite is true, of course, if
v< 0.
In Fig. l.lO, the voltage v may be time varying, or it may be constant, in which
case we would probably label it V. Another symbol that is often used for a constant
voltage source, such as a battery with V volts across
terminals, is shown in Fig. its 1 . 1 1
6
6b 6b
2A 2A
+ +
12 V 12 V
~
o o
(0) (b)
FIGURE 1.1 2 Independent current FIG U R E 1 .1 3 (a) A Source delivering and (b) Absorb-
source ing power
—
12 Introduction Chap. 1
battery is delivering 24 W to the external circuit. In Fig. 1 . 1 3(b) the battery is absorbing
24 W, would be the case when it is being charged.
as
The sources that we have discussed here, as well as the circuit elements to be
considered later, are ideal elements. That is, they are mathematical models that approxi-
mate the actual or physical elements only under certain conditions. For example, an
ideal automobile battery supplies a constant 12 V, no matter what external circuit is
connected to it. Since its current is completely arbitrary, it could theoretically deliver
an infinite amount of power. This, of course, is not possible in the case of an actual
device. A real 12-V automobile battery supplies approximately constant voltage only
as long as the current it delivers is low. When the current exceeds a few hundred
amperes, the voltage drops appreciably from 12 V.
We shall consider practical sources in a later chapter and see under what conditions
they may be approximated by ideal sources. Also later, we shall consider dependent
sources whose voltage (or current) is controlled by another voltage or current some-
where else in the circuit.
EXERCISE
1.4.1 Find the power being supplied by the sources shown.
Ans. (a) 36 W, (b) -40 W, (c) -2 W
IA
t
'
o
-2V
EXERCISE 1.4.1
Let us now look at the words circuit analysis, which are contained in the title of the
book, and see what they mean. Generally, if an electric circuit is subjected to an input
or excitation in the form of, say, a voltage or a current provided by an independent
source, then an output or response is produced. The output or response may also be a
some element in the circuit. There may be, of
voltage or a current associated with
course, more than one input and more than one output.
There are two main branches of circuit theory, and they are derived from the
following three key words: input, output, and circuit. The first branch is circuit
Chap. 1 Introduction 13
analysis, which, given the circuit and the input, is concerned with finding the output.
The other branch is circuit synthesis, which, given the input and output, is concerned
with finding the circuit itself.
Network synthesis is much more complex in general than analysis and will proba-
bly be encountered by the student in a later course. Circuit analysis will be our concern
in this book. We may be interested only in finding one or more outputs, such as a
voltage or current existing somewhere in the circuit, or in determining the energy or
power delivered one element or another. Or we may wish to perform a complete
to
analysis, finding every unknown current and voltage in the circuit. In any case, in the
chapters to come we shall develop systematic methods of analysis that can be applied
generally to any circuit of the type we consider.
PROBLEMS
1.1 Let f(t) in the graph shown be the charge q{t) in millicoulombs that has entered
the positive terminal of an element as a function of time, (a) Find the total charge
that has entered the element between 3 and 7 s, (b) the current entering the element
at 5 s, and (c) the energy absorbed by the element between 1 and 7 s if the voltage
across the element is 10 V.
f(t)
4 -
T (seconds)
PROBLEM 1.1
1.2 Let/(/) in Prob. 1.1 be the current i(t) in milliamperes entering an element ter-
minal as a function of time, (a)Find the charge that enters the terminal between
and 7 s and (b) the rate at which the charge is entering at t = 5 s.
1.3 The terminal voltage of a voltage source is given by v = 10 cos 100/ V. If the
charge leaving the positive terminal is q = 6 sin 100/ mC, find the power being
supplied by the source at t = 71/100 s.
1.4 For the element shown, the charge entering the positive terminal is
q = —2e' 05 '
C. Find the power delivered to the element as a function of time
and the total energy delivered to the element between and 2 s if (a) v = 6/',
o-
+
PROBLEM 1.4
1.5 If the element in Prob. 1.4 is a battery with a constant terminal voltage of v = 12 V
and the current i is 2 A, then the battery is in the process of being charged. (It is
absorbing rather than delivering power.) (a) Find the energy supplied to the bat-
tery in 2 h (hours), (b) Find the charge delivered to the battery in this time. Note
the consistency of the units, 1 V = 1 J/C.
1.6 Suppose the voltage v in Prob. 1.5 varies linearly from 4 to 12 V during the 2 h
that the battery is being charged. If / = 2 A is constant during this period, find
(a) the total energy supplied and (b) the total charge delivered to the battery.
1.7 In the element of Prob. 1 .4, let i(t) = for t < and i(t) = 6 sin It A for t > 0.
If v = 4 dijdt V, show that the energy delivered to the element is nonnegative for
all t > 0.
1.8 In Prob. 1 .7, find the total charge delivered to the element and the power absorbed
at / = 7r/4 s.
1-*'"7
2.
2
RESISTIVE CIRCUITS
The simplestand most commonly used circuit element is the resistor. All electrical
conductors exhibit properties which are characteristic of a resistor. When currents
flow in conductors, electrons which make up the current collide with the lattice of
atoms in the conductor. This, of course, on the average, impedes the motion of the
electrons. The larger the number of collisions, the greater the resistance of the con-
ductor. We shall consider a resistor to be any device which exhibits solely a resistance.
Materials which are commonly used in fabricating resistors include metallic alloys and
carbon compounds.
In this chapter, we shall first introduce the terminal relations for a resistor based on
Ohm's law. Two laws necessary for systematic solutions of networks, known as
Kirchhoff's laws, are then examined. With these laws, we shall begin our study of
by finding solutions for "single-loop" and "single-node-pair" resistive
circuit analysis
networks having independent sources as inputs. We shall conclude the chapter with a
discussion of simple measuring instruments followed by a discussion of practical
resistors.
Georg Simon Ohm (1787-1854), a German physicist, is credited with formulating the
current-voltage relationship for a resistor based on experiments performed in 1826. In
1827 he published the results in a paper titled "The Galvanic Chain, Mathematically
Treated." As a result of this work, the unit of resistance is called the ohm. It is ironic,
however, that Henry Cavendish (1731-1810), a British chemist, discovered the same
results 46 years earlier. Had he not failed to publish his findings, the unit of resistance
might well be known as the caven.
15
16 Resistive Circuits Chap. 2
Ohm's law states that the voltage across a resistor is directly proportional to the
current flowing through the resistor. The constant of proportionality is the resistance
value of the resistor in ohms. The
symbol for the resistor is shown in Fig. 2.1.
circuit
For the current and voltage shown, Ohm's law is
m = m) (2.i)
where R >
is the resistance in ohms.
1 ohm = 1 V/A
The symbol used to represent the ohm is the capital Greek omega (Q).
Since R is constant, (2.1) is the equation of a straight line. For this reason, the
resistor is called a linear resistor. A graph of v(t) versus i(t) is shown in Fig. 2.2, which
is a line passing through the origin with a slope of R. Obviously, a straight line is the
only graph possible for which the ratio of v{t) to /(/) is constant for all /'(/).
v(t)
i(t)
+
Ht)
v(t)
FIGURE 2.1 Circuit FIGURE 2.2 Voltage-current FIGURE 2.3 Typical voltage -
symbol for the resistor characteristic for a linear resistor current characteristic for a non-
linear resistor
In reality, all practical resistors are nonlinear because the electrical characteristics
of all conductors are affected by environmental factors such as temperature. Many
materials, however, closely approximate an ideal linear resistor over a desired operating
region. We shall concentrate on these types of elements and simply refer to them as
resistors.
An examination of (2.1), in conjunction with Fig. 2.1, shows if /'(/) > (current
entering the upper terminal), then v(t) > 0. Thus the current enters the terminal of
higher potential and exits from that of the lower potential. Next, suppose that/(r) <
7
(current entering the lower terminal). Then v(t) < 0, and the lower terminal is higher in
potential than the upper one. Once again, the current enters the terminal of higher
potential. Since charges are transported from a higher to a lower potential in passing
through the resistor, the energy lost by a charge </ [energy = qv(t)] is absorbed by the
resistor in the form of heat. The rate at which energy is dissipated is, by definition, the
instantaneous power
(2.2)
A graph of (2.2), shown in Fig. 2.4, reveals that p{t) is a parabolic (and thus
nonlinear) function of i{t) or v{t) which is always positive. (The horizontal scales are,
of course, different in the two cases.) Thus, for a linear resistor, the instantaneous
power is nonlinear even though the voltage-current relationship is linear.
Therefore, since p(t) is always positive, we see that the above integral is positive and
that the resistor is, indeed, a passive element.
The beginning student often encounters difficulty in determining the proper
algebraic sign in applying (2.1) when the voltage assignment differs from that of Fig.
2.1. Consider, for example, the assignment of Fig. 2.5. Comparing these figures, we see
that t>,(?) = — v(t) = —Ri(t). Therefore, when the voltage assignment is such that the
assumed current enters the terminal of lower potential (— ), a minus sign must be
employed in using (2.1). It should be noted that
„(a
py)
_ 1& - r-*>i(')]
2
_ vM
R R R
G =1 (2.3)
The unit of conductance is the mho (A/V), which is ohm spelled backwards. The
symbol representing the mho is an inverted omega (13). Combining (2.1)— (2.3), we see
that alternative expressions for Ohm's law and instantaneous power are
and
As an example of the application of Ohm's law, consider finding the current for the
circuit of Fig. 2.6 in which a 12-V storage battery is connected to a 1-kQ resistor.
From (2.3) and (2.4), G = j^ = 10" 3 yand/ = 10" 3 x 12 A= 12 mA. Also, (2.5)
yields p(t) = 10" 3 x 12 2 = 144 mW, which is the minimum power rating or wattage
required for the resistor to ensure that it will not be damaged as a result of over-
1
heating.
p(t)
i(t)
-f
FIGURE 2.4 Graph of the instan- FIGURE 2.5 Resistor F I GURE 2 . 6 Simple resistive
taneous power for a resistor with a reversed voltage circuit
assignment
The current in this example is a direct current since its value is invariant in time.
Suppose we now replace the 12-V battery by the time-varying voltage v{t) = 10 cos t V
and repeat the above procedure. The current is
l0
i(t) =
^ 1
A= 10 cos t mA
and the instantaneous power is
EXERCISES
2.1.1 The terminal current of a 20-kQ resistor is 10 mA. Find (a) the conductance,
(b) the terminal voltage, and (c) the minimum wattage of the resistor.
Arts, (a) 5 x 10" 5 15, (b) 200 V, (c) 2 W
2.1.2 The instantaneous power absorbed by a 1-kQ resistor is 10 sin 2 377/ W. Find
v(t) and i(t). Am. 100 sin 377/ V, 0.1 sin 377/ A
J
The wattage of a resistor is based on the average power (to be discussed in Chapter 12). For
direct currents, the instantaneous power equals the average power.
9
5A
EXERCISE 2.1.3
The solution of single-resistor circuits, such as those in the previous examples, may be
found using Ohm's law; however, more complicated circuits require the use of two
laws first stated by the German physicist Gustav KirchhoiT(l 824-1887) in 1847. 2 The
two laws are formally known as Kirchhoff 's current law and Kirchhoff 's voltage law.
These laws, together with the terminal characteristics for the various circuit elements,
permit systematic methods of solutions for any electrical network. We shall not attempt
to prove Kirchhoff *s laws here since the concepts necessary for the proof are developed
in a later, interesting study of electromagnetic field theory.
A circuit consists of two or more circuit elements connected by means of perfect
conductors. Perfect conductors are zero-resistance wires which allow current to flow
freely but accumulate no charge and no energy. In this case, the energy can be con-
sidered to reside, or be lumped, entirely within each circuit element, and thus the
network is called a lumped-parameter circuit.
A point of connection of two or more circuit elements is called a node. An example
of a circuit with three nodes is shown in Fig. 2.7(a). Node 1 consists of the entire con-
Node i
Node I
Node 3^
© Node 3
Node 2
Node 2
(o) (b)
nection at the top of the circuit. The beginner quite often mistakes points a and b for
nodes. It should be noticed, however, that a and b are connected by a perfect con-
ductor and can be considered electrically as being identical points. This is readily
demonstrated by redrawing the circuit in thewhere at node 1 all form of Fig. 2.7(b),
connections are shown at a single point. Similar comments apply for node 2. Node 3 is
required for the interconnection of the independent voltage source and the resistor.
With these concepts, we are now ready to discuss the all-important laws of Kirchhoff.
Kirchhoff's current law (KCL) states that
To demonstrate the use of this law, consider currents flowing into a node, as shown in
Fig. 2.8. KCL states that
jl + 2 + (-/s) + i4
= <L
where we recall that i3 flowing out of the node is equivalent to — i3 entering the node.
For the sake of argument, let us suppose that the sum is not zero. In such a case, we
would have
U +h- /s + /« - * 9* o
where and hence must be the rate at which charges are accumulated
*P has units of C/s
in the node. However, a node consists of perfect conductors and cannot accumulate
charges. In addition, a basic principle of physics states that charges can neither be
created nor destroyed (conservation of charge). Therefore, our assumption is not
valid, and ¥ must be zero, demonstrating the plausibility of KCL.
Suppose in our example we now consider the sum of the currents leaving the node.
From Fig. 2.8, the sum is
—/, - i2 + i3 —U =
or, multiplying both sides by —1, we see that
i, + i2 — i3 + i4 =
which is identical to our previous result. This demonstrates an equivalent statement for
KCL, which states that
^^K^r + i2 *
4 =h
where /,, i 2 , and /4 are entering the node and i3 is leaving. This form of the equation
illustrates another statement for KCL, stated as
The sum of the currents entering any node equals the sum of the currents leaving the
node.
Chap. 2 Resistive Circuits 21
N
E =o /. (2.6)
where i„ is the nth current entering (or leaving) the node and N is the number of node
currents.
As an example of KCL. let us find the current i in Fig. 2.9. Summing the currents
entering the node, we have
5 + -
/ (-3) -2 =
or
i = -6 A
FIGURE 2.8 Currents flowing into a node FIGURE 2.9 Example of KCL
AWe = —6 A
note that / entering the node is equivalent to 6 A leaving the node.
Therefore it is not necessary to guess the correct current direction prior to solving the
problem. We still arrive at the correct answer in the end.
If we sum the currents leaving the node, we have
-5-/+ (-3) + 2 =
or
/= -6 A
As an
The algebraic sum of the voltages around any closed path is zero.
(&Q
illustration, application of this statement to the closed path abcda offFii
Fig. 2.10
gives
where the algebraic sign for each voltage has been taken positive when going from +
to — (higher to lower potential) and negative when going from — to + (lower to
higher potential) in traversing the element.
As in the case of KCL, we shall not attempt a proof of KVL. However, to illustrate
the plausibility of (2.7), let us assume that its right member is not zero. That is,
—v + x
v2 — v3 = <S> ^
The left member of this by definition the work required to move a unit
equation is
traversed. Consider, for example, the path adcba in Fig. 2.10. Summing the voltages, we
find
V3 — V2 + *>1 = °
where v„ is the nth voltage in a loop of N voltages. The sign of each voltage is chosen
as described earlier for (2.7).
As an example of the use of KVL, consider finding v in Fig. 2.1 1. Traversing the
circuit in a clockwise direction, we find
1 - [r^
+ v -
—V/r-
T
+W ^
^AA/
a a + 2V -
FIGURE 2.10 Voltages around a closed path FIGURE 2.11 Circuit to illustrate KVL
(r
Chap. 2 Resistive Circuits 23
5 + 2-10-v =
or v = —3 V. which, of course, is the same result obtained for the clockwise traversal.
Still another version of KVL for Fig. 2.1 1 yields
10 +v=2+3
where the sum of the voltages with one polarity is equated to the sum of the voltages
with the opposite polarity. (Stated another way, the voltage rises equal the voltage
drops.)
In each of the previous examples, KVL has been applied around conducting paths,
such as abcda above. The law, however, is valid for any closed path. Consider, for
instance, the path acda of Fig. 2.11. We note that movement directly from a to c is not
along a conducting path. Applying KVL to this closed path yields v ac + 10 — 2 = 0,
where vac is the potential of point a with respect to c. Thus, v ae = — 8 V. We could
also have chosen the path abca, for which
-5 + v — vac = —5 - 3 - v ac =
Therefore. vac = — 8 V. which demonstrates the use of different closed paths to
obtain the same result.
As another example of the application of KCL
and KVL, consider finding i x and
vx in the network of Fig. 2.12. Summing
node a gives
the currents entering
—4 — 1 + /, = 0, or i, = 3 A. At node b, — /, + 2 — i 2 = 0, or i 2 = — 1 A. At
node c, i 2 + i 3 — 3 = 0, or i 3 = 4 A. Therefore, at node d, —i x — 1 — i 3 = 0, or
i x = —5 A. Next, KVL about the path abcda gives — 10 + v 2 — v x = 0. From Ohm's
law, t\ = 5/ 2 = -5 V. Therefore, v x = — 15 V.
Before concluding our discussion of Kirchhoff's laws, consider the network of
Fig. 2.13. in which several elements are shown within a closed surface S. We recall
that the current entering each element equals that leaving the device so that each
element stores zero net charge. Therefore, the total net charge stored within the
surface is zero, requiring that
h +h+ ii +U=
This result illustrates a generalization of KCL, which states
The algebraic sum of the currents entering any closed surface is zero. 3
To illustrate the plausibility of the generalized KCL, let us write KCL equations at
nodes a, b. c, and d of Fig. 2.13. The results are
3
The surface cannot pass through an element which is considered to be concentrated at a point
in lumped-parameter circuits.
24 Resistive Circuits Chap. 2
SS\\>5 \ .CLOSED
§f/ \V \^ SURFACE S
FIGURE 2.12 Network for example of KCL FIGURE 2.13 Network for generalized KCL
and KVL
'
= — l\ if.
—h -I*
= —h + it
m = h+h
Adding these equations yields
EXERCISES
2.2.1 Identify the nodes in the figure.
ion
£> -AAA/—
i
Vw-
2fl
3.5 A 10 a (t) 2A
•"O
3A
-AAA^
IA
EXERCISE 2.2.1
Chap. 2 Resistive Circuits 25
2.2.2 Determine (a) /,. (b) /, . and (c) the potential of a with respect to h.
EXERCISE 2.2.2
Elements ar e said to be connected in ser ie s when they al l ca rry the same current.
Clearly, the networks of this section consist entirely of elements connected in series.
An important circuit of this type, consisting of two resistors and an independent
voltage source, provides an excellent starting point. We shall first analyze this special
case and then develop the more general case.
A single-loop circuit having two resistors and an independent voltage source is
shown in Fig. 2.14(a). The first step in the analysis procedure is the assignment of
currents and voltages to all elements in the network. In this circuit, it is obvious from
KCL that all elements carry the same current. We may arbitrarily call this current / in
the direction shown (clockwise). The novice often attempts to guess the true current
direction in making assignments. Such an assignment is not necessary, and is not
usually possible, even for the expert. We next make the voltage assignments for i?j and
R 2 as i\ and r 2 respectively. These assignments are also arbitrary but in the figure have
,
(a) (b)
The second step in the analysis is the application of KVL, which yields
v = Wj + v2
where, from Ohm's law,
t>, = A\/
(2.9)
R 2i
v = R^i +R 2
i
(2.10)
R, + R2
Let us now consider a simple circuit consisting of the voltage source v connected to
a resistance R s , as shown in Fig. 2.14(b). If R s
is selected such that
v
1
> = R,
(2.11)
then the network is called an equivalent circuit of Fig. 2.14(a) because an identical
current (response) is produced for a voltage v (excitation). In general, two circuits are
said to be equivalent when they exhibit identical voltage-current relationships at their
terminals.
Comparing (2.10) and (2.11), it is obvious that
R = s R, +R 2 (2.12)
R,
v, =
R + R2
l
(2.13)
R*
R, A'-
Chap. 2 Resistive Circuits 27
The potential of source r divides between resistances /?, and /?, in direct proportion to
their resistances, demonstrating the principle of voltage division for two series resistors.
We s ee in (2. 1 3) that the greater voltage appears across the larger resistor,
I he instantaneous powers absorbed by /?, and R 2
are
Pl
Pi = i^
R x
(R.
*<
+ R.Y
*>
1 V=*R
and
_v\
~ _ R2 V 2
Pl
R -(R + R y2 i 2
p ^ p> = RT^r
v2
l
I
iR7+Rj
v \
=
== VI
The p ower delivered by the source also equals vi. indicating that the powe r delivered by
R and R 2 This result is
tne source eq uals tha t absorbed by t
. known as conservation oj
power (sometimes also referred to as Tellegen s theorem), a property which is often
useful in circuit analysis.
Let us now digress briefly and repeat the analysis for the counterclockwise current
assignment of Fig. 2.15. Application of KVL yields
V = vl + vz
in which
»i = ~ R xh
v2 = -R 2i x
v = -RJ, - R 7ix
from which
i, =
R ^-R
t 2
90
72 sin t 120 sin t
90 + R,
sin 2 1 W and p 2 = R 2 i 2 = 38.4 sin 2 t W. Thus, the power delivered by the source is
96 sin 2 / W. We also observe that the power absorbed by R and R 2 should equal that x
v l|R|
6 V
2|R 2
+ '
n|R n
v
FIGURE 2.15 Single-loop circuit with FIGURE 2.16 Single-loop circuit with
Let us now extend our analysis to include the series connection of A^ resistors and
an independent voltage source, as shown in Fig. 2. 16. KVL gives
v = v. + vN
in which
v2 =R 2i
(2.14)
vN = RN i
Therefore
v = R,i +R + 2i . . . + RN i
v
i = (2.15)
/?, +R 2 R*
Si s
Chap. 2 Resistive Circuits 29
Therefore the equivalent resistance of N series resistors is simply the sum of the
individual resistances.
Substituting (2.15) and (2.16) into (2.14). we find
— R,
tllv
R,
(2.17)
Rv
which are the equations describing the voltage division property for TV series resistors.
Again, we see that the voltage divides in direct proportion to the resistance.
The instantaneous power delivered to the series combination, from (2.2) and
(2.17). is
vl
' = # + #+•••+*
R R, 2
~ R:
t
9
v + . + ^v 2
R] Ri
= VI
R.
This power is equal to that delivered by the source, verifying conservation of power for
the series connection of TV resistors.
EXERCISES
2.3.1 For the network shown, find (a) an equivalent circuit, (b) the current (c) the ;',
power delivered by the source, (d) v u and (e) the minimum wattage required
for /?!. Am. (b) 0.5 A, (c) 5 W, (d) -4 V, (e) 2 W
ion
EXERCISE 2.3.1
30 Resistive Circuits Chap. 2
2.3.2 In Fig. 2.14 (a), v = lOe"', v 2 = 2e~', and /?, = 16 0. Find (a) R2 ,
(b) the
instantaneous power delivered to R 2 and (c) , the current i.
available to operate the load. Referring to Fig. 2.14 (a), if R2 represents the load
and v the 12-V battery, find (a) the current i, (b) the necessary resistance R x ,
first assign voltages and currents to each circuit element. As in the case of the single-
loop circuit, the assignments are completely arbitrary. We have chosen the assign-
ments given by /,, i 2 , and v.
(a) (b)
Next, we need to apply either KCL or KVL. Inspection of the circuit shows that a
single node (either the upper or the lower) is common to all elements. This suggests
that for single-node-pair circuits KCL is the most efficient choice. Applying KCL at
the upper node yields
/, = = G<o
(2.18)
h = = G2 v
Combining these equations gives
i =Gv+G 2v
4 A load is an element or collection of elements connected between the output terminals. In this
v = (2I9 >
Grnr t
vp ~ = ±- = v (2.20)
Gp
then the network is an equivalent circuit to that of Fig. 2.17(a). Comparing (2.19) and
(2.20). we see that
GP = G + G ] Z (2.21)
or
(2.22)
G, .
(2.23)
G, .
'i-g—tg;'
R2 .
(2 24)
'
A,
Therefore the current divides in inverse proportion to the resistances. We see that the
larger current flows through the smaller resistor. The power absorbed by the parallel
32 Res/stive Circuits Chap. 2
combination is
Pi — R-lh + *2*2
R\i
2^1 + +R 2^2
(*1 + *2)
2
(#. 2)
*1*2 :2
VI
R, +R 2
i = /, + i2
for which
/, = G,v
(2.25)
i N = GN v
Therefore we have
i = G{o +G 2
v + . . . + GN v
from which
v — (2.26)
G,+G + 2 ... +G t
If we now select G in Fig. 2.17(b) such that (2.20) is satisfied, then (2.26) requires
that
Gp = G, +G + 2 . . . + GN = £ i = i
G, (2.27)
R
- -+-
R +
/?,
... +- =£
R N i=i
-
Rj
(2.28)
2
-
Hence the reciprocal of the equivalent resistance is simply the sum of the reciprocals
of the resistances.
Combining (2.25)-(2.28), we find
G,. Rp .
R,,
R,
(2.29)
. _ Gjv .
_ /?„ .
R, 12
= TU
and Rp = 2 Q.
Let us now consider finding the equivalent resistance R eq of the network of Fig.
2.19(a). as viewed from terminals x-y. Such reductions are very helpful in analyzing
many types of circuits, as we shall see in the next section. The process is carried out by
7a i n 7 A
*eq^ 12 A 5 n 6 A
(a) (b)
7 A
*o- ^AA/—
:4 a R eq — II A
(c) (d)
successive combinations of parallel and series connected resistors. In Fig. 2.19(a), the
student often errs in taking combinations such as the 7- and 12-fi resistors to be in
series. We see, at node a, a current in the 7-Q resistor would divide
however, that
between the and 12-Q resistor; hence they cannot be in series. The 1- and 5-fi
1-
resistors, however, would carry the same current. Therefore they are in series, having a
series resistance of 6 Q as shown in Fig. 2.19(b). We now observe that the same voltage
would occur across the 6- and 12-Q resistors, indicating a parallel connection having
an equivalent resistance of (6)(12)/(6 + 12) = 4Q, as shown in Fig. 2.19(c). It is
apparent that the 7- and 4-fi resistors of this network are in series, yielding an equiva-
lent resistance for the entire network of 1 Q [Fig. 2.19(d)]. Therefore, from terminals
1
x-y, the network could be replaced by a single resistor of 1 1 Q. This is useful in deter-
mining, for instance, the power delivered by a source connected to terminals x-y.
Suppose a 22-V source is applied. Then the current flowing from the source is / = ^
= 2A, which gives an instantaneous power p(t) = (22)(2) = 44 delivered to the W
resistor network.
EXERCISES
2.4.1 In the circuit shown, find (a) the equivalent resistance seen from the source
terminals, (b) v u (c) i2 , and (d) i.
EXERCISE 2.4.1
(b) i, and (c) i} . Ans. (a) 40 Q, (b) 0.5 sin t A, (c) 0.25 sin / A
2.4.3 A load requires 4 A and absorbs 16 W. A current source of 10 A is available.
For the circuit of Fig. 2.17 (a), if G 2 and i the current source,
represents the load
find (a) the voltage v and (b) the required conductance C,.
Ans. (a) 4 V, (b) 1.5 U
2.4.4 Find the equivalent resistance, R eq . Ans. 9.4 Q
Chap. 2 Resistive Circuits 35
v
eq
EXERCISE 2.4.4
For our first example, let us consider a single-loop circuit containing three resistors and
two independent voltage sources, as shown in Fig. 2.20(a). KVL and Ohm's law give
10 + 100/ =
for which = -0.1 A and y, = 30/ = — 3 V.
i
We see that the reduced equation is satisfied by the equivalent circuit of Fig. 2.20(b),
where the voltage sources are replaced by sum and the resistances by
their algebraic
their series equivalent. From this circuit, power absorbed by all resistors is
the
/'loon
= 100/ 2 = 1 W. Therefore the net power delivered by the two voltage sources
must be 1 W. We recall that the power absorbed by an element is the product of the
voltage and the current, where the current enters the positive voltage terminal. For the
10-V source, we see that /> 10V = (10)(-/) = (10)(0.1) = 1 W absorbed. Physically,
this means that current is flowing into the positive terminal and that the source
is receiving power, or is being charged. The absorbed power in the 20-V source is
p 2Q v = 20/ = (20)( — 0.1) = — 2 W. The minus sign indicates that the source is
delivering 2 W to the circuit. Hence the net power from the two sources is 1 W, which
is the power delivered to the resistors. Finally, let us find the potential of point a with
20 A 30 n,
'VW
+ v, -
,.v(*)
so a
© 20 V I ioo a
-w\, —
(a) (b)
+
10 s n*t(T) v(t)§.OI U 02 V 05A 07 U
0.1 15 (a 10-fi resistor) would be an equivalent circuit for the network [see Fig. 2.17(b)].
The total instantaneous power absorbed by the conductances is therefore
= (10 sin nt - 5)
2
Pab;
0.1
p x
= 10 sin nt(\00 sin nt - 50) W
Similarly, the 5-A source delivers
As a third example, consider finding r. the power delivered by the source, and
/.
thepower absorbed by the 8-fi load of Fig. 2.22(a). We begin by obtaining successive
combinations of parallel and series resistor connections. The 4- and 8-fi resistances
(in series) add to give 12 fi. These 12 Q are in parallel with the 6-fi resistor, giving an
equivalent value of (12)(6)/(12 -4- 6) = 4fi [Fig. 2.22(b)]. We now add the 12- and
4-fi resistances, which are in parallel with the 16 Q, giving (16)(16)/(16 + 16) = 8 Q,
as shown in Fig. 2.22(c). This is the equivalent resistance as seen looking into the
circuit at .v-.v. Applying KVL to the simplified circuit, we have
-30e" 2/ + 8/ =
from which
i = 3<r 2 '
A
Vl = (-Jr^30<r = 2'
24<r 2 'V
(o)
30e" 30e
lb) (c)
which is the voltage across points x-x in the circuit. Proceeding to Fig. 2.22(b), we see
that v t
is the voltage across the series combination of the 12- and 4-fi resistors;
—
U2 = t,1=6e
" 2'
v
(T2T4)
which is the voltage across points y-y in the circuit. In Fig. 2.22(a), it is obvious that
vz is the voltage across the series connection of the 4- and 8-fi resistors. Therefore
voltage division requires that
4e~ 2 'V
(wh) v >
p2 = V = 2e-*' W
As a final example, let us find the current i in Fig. 2.23(a). We first combine the two
6-Q resistors on the right of Fig. 2.23(a) and obtain the equivalent resistance which is
in parallel with the 4-fi resistor. The result, (4)(12)/(4 + 12) = 3 Q, is shown in
Fig. 2.23(b). If we now replace the two series resistors to the right of points x-x and
the parallel 3- and 6-fi resistors to the left of x-x by their respective series and parallel
equivalents, we obtain the circuit of Fig. 2.23(c). Using current division, we find
x12 = 3a
'HtT6)
A second application of current division to Fig. 2.23(a) reveals immediately that
x3 =
'-(fTTT,)'.- (1) TA
x 3n 6 A
H *AV ^A\
6A
(b) (c)
EXERCISES
2.5.1 Find v, r ab , and the power delivered by the 5-V source.
Ans. -10 V, -8.33 V, -0.167 W
20.fl 60A
-• "vVv-
30 A b 40 a
EXERCISE 2.5.1
2.5.2 Find G and construct an equivalent circuit having one current source and a
single conductance. Ans. G = 0.03 15, i = 3 sin t A directed upward
2 sin t A
(| J
5 .01 IT |.02 V () 5 sin t A G550sintV
EXERCISE 2.5.2
2.5.3 Find the power absorbed by the 90-Q resistor. Ans. 3.6 W
24n.
EXERCISE 2.5.3
2A 4n: 6A 3A :6fi 2n
EXERCISE 2.5.4
40 Resistive Circuits Chap. 2
terminals and has zero voltage across its terminals. In contrast, an ideal voltmeter
measures the voltage across its terminals and has a terminal current of zero. An ideal
ohmmeter measures the resistance connected between its terminals and delivers zero
power to the resistance.
The practical measuring instruments that we shall consider only approximate the
ideal devices. The ammeters, for instance, will not have zero terminal voltages.
Likewise the voltmeters will not have zero terminal currents, and the ohmmeters will
not have zero power delivered from their terminals.
A popular type of ammeter consists of a mechanical movement known as a
D'Arsonval meter. This device is constructed by suspending an electrical coil between
the poles of a permanent magnet. A dc current passing through the coil causes a
rotation of the coil, as a result of magnetic forces, that is proportional to the current.
A pointer is attached to the coil so that the rotation, or meter deflection, can be
visually observed. D'Arsonval meters are characterized by their full-scale current,
which is the current that will cause the meter to read its greatest value. Meter move-
ments are common having from 10 juA to 10 mA.
full-scale currents
series with a resistance RM , as shown in Fig. 2.24. In this circuit, RM represents the
resistance of the electrical coil. Clearly, a voltage appears across the ammeter terminals
as a result of the current i flowing through RM RM
. is usually a few ohms, and the
terminal voltage for a full-scale current is nominally from 20 to 200 mV.
The D'Arsonval meter of Fig. 2.24 is an ammeter which is suitable for measuring
dc currents not greater than the full-scale current 7FS Suppose we wish, however, to
.
measure a current which exceeds /Fs It is apparent that we must not allow a current
.
greater than 7FS to flow through the device. A circuit to accomplish this is shown in
Fig. 2.25, where R p is a parallel resistance which reduces the current flowing through
the meter coil.
* -i M
"(D
'
<^>
FIGURE 2.24 Equivalent circuit for a FIGURE 2.25 Ammeter circuit
D'Arsonval meter
Chap. 2 Resistive Circuits 41
R,
* FS
— 'fs
R. R
where iFS is the current which produces /, s in the D'Arsonval meter. (Clearly, this is
(2.30)
p i
'FS
— YI F
scale voltage, v = v FS , occurs when the meter current is 7FS Therefore, from
. KVL,
RJ
s'FS R.uhs —
1(2
from which
R — 7^ — Rm
s
t
V (2.31)
The current sensitivity of a voltmeter, expressed in ohms per volt, is the value
obtained by dividing the resistance of the voltmeter by its full-scale voltage. Therefore
Q/V rating = R. + R, R
(2.32)
-E + (R, + RM + R x - )i
from which
R, (Rs + RM )
-SVW-
R< + R,
R* =
(¥ ~ Y +
R l
* Rm) (2.33)
EXERCISES
scale voltage of 100 V using a D'Arsonval meter with (a) R M = 100 fi and
7FS = 50 fiA and (b) R M = 50 fi and 7FS = 1 mA.
Arts, (a) 2 Mfi, 20 kfi/V, (b) 100 kfi, 1 kfi/V
2.6.3 What voltage would each meter design of Ex. 2.6.2 measure in the circuit
below? Why are the two measurements different?
Am. 99.5 V, 90.9 V
10 mA
EXERCISE 2.6.3
2.6.4 The meter movement of Ex. 2.6.1 is used to form the ohmmeter circuit of Fig.
2.27. Determine R s and E so that /'
= / FS /2 mA when R x = 10 kfi.
Am. 9.95 kfi, 10 V
.
Resistors are manufactured from a variety of materials and are available in many
sizesand values. Their characteristics include a nominal resistance value, an accuracy
with which the actual resistance approximates the nominal value (known as tolerance).
a power dissipation, and a stability as function of temperature, humidity, and other
;i
environmental factors.
The most common type of found in electrical circuits is the carbon com-
resistor
position or carbon film resistor. The composition type is made of hot-pressed carbon
granules. The carbon film device consists of carbon powder which is deposited on an
insulating substrate. A typical resistor of this type is shown in Fig. 2.28. Multicolored
% Tolerance
4)iiD) y
a b c
°
bands, shown as a. b, c. and on the resistor body to indicate
tolerance, are painted
the nominal value of the resistance. The color code
bands is given in Table 2.
for the 1
Bands a. b. and c give the nominal resistance of the resistor, and the tolerance band
gives the percent that the resistor may deviate from the nominal value. Referring to
Fig. 2.28, the resistance is
Bands a, b, and c
Gold* -2 Yellow 4
Silver* -1 Green 5
Black Blue 6
Brown 1 Violet 7
Red 2 Gray 8
Orange 3 White 9
% Toleranc e Band
Gold ±5%
Silver ±10%
R^(\o* (- b) \b
c\-
44 Resistive Circuits Chap. 2
As an example, suppose we have a resistor with band colors of yellow, violet, red,
and silver. The resistor will have a value given by
R= (4 x 10 + 7) x 10 2 ± 10%
= 4700 ± 470 Q
RESISTIVE
LAYER
EXERCISE
2.7.1 Find the resistance range of carbon resistors having color bands of (a) brown,
black, red, silver; (b) red, violet, yellow, silver; and (c) blue, gray, silver, gold.
Ans. (a) 900-1 100 Q, (b) 243-297 kQ, (c) 6.46-7.14 Q
PROBLEMS
2.1 A 5-V battery is connected to the ends of a 1000-ft length of conducting wire
and a 10- A current flows. What is the resistance per foot of the wire?
2.2 A 2-kQ resistor is connected to a battery and 10-mA flow. What current will
flow if the battery isconnected to a 500-Q resistor?
2.3 A 200-Q resistor is connected to a 12-V source. Find the current and the mini-
mum wattage of the resistor. If a 100-fi resistor is inserted in series in the circuit,
what is the voltage across it ?
PROBLEM 2.4
2.5 Several series resistors draw 100 mA from a 10-V source. If a 1-kfi resistor is
2.6 Two resistors connected in series draw 25 mA from a 100-V source. The voltage
across one of the resistors is 25 V. Find the value of each resistor.
46 Resistive Circuits Chap. 2
2.7 Design a voltage divider to provide 1, 5, 10 and 20 V from a 25-V source. The
source is to deliver 25 mW of power to the divider.
2.8 A 1-A current source is connected to the parallel combination of 20- and 30-Q
resistors. Find the voltage, current, and minimum wattage of each resistor.
2.9 A 6-Q resistor is added in parallel with the combination of Prob. 2.8. Find the
voltage and current for this resistor.
2.11 A 4-fi and a 12-fi resistor are in parallel. The parallel combination is in
resistor
series with a 7-Q resistor and a voltage source of 10 sin It V. Find the current
in the 4-Q resistor and the instantaneous power from the source.
2.12 Two 1-kfi resistors are in series. When a resistor R is connected in parallel with
one of them, the resistance of the combination is 1750 Q. Find R.
2.13 Find R ea .
ea
6fl
l
eq irx> 2 n
-AAAr-
5A 7 A
PROBLEM 2.13
2.15 A current source is connected to the network of Prob. 2.13 and i = 3 cos t A.
What is the value of the current source?
2.16 Find i.
PROBLEM 2.16
2.17 Repeat Prob. 2.16 with the 6-V source replaced by a current source of 6 A.
2.18 Find R.
Chap. 2 Resistive Circuits 47
4a 5n
-'Wv/-
2A
20 V
© 3flS
PROBLEM 2.18
I2A
12 V
PROBLEM 2.19
2.21 Find R.
t V>A—
+ 30V
+
I5A( ii
3A< 4A v \ i2n
PROBLEM 2.21
PROBLEM 2.23
48 Resistive Circuits Chap. 2
2.24 Find v(t) and the instantaneous power to the 3-ft resistor.
+
v(t)> in.
© I A
f
2a
G) e
"A 3A
PROBLEM 2.24
4 n
4 A 4 A
-^\AA/ O-
+
4 A
-© V(t) Q4s,n t V
-AAA, O-
4A 4A
PROBLEM 2.25
2.27 A 20,000-Q/V voltmeter has a full-scale voltage of 120 V. What current flows in
the meter when measuring 90 V?
2.28 Two 10-kft resistors are connected in series across a 100-V source. What voltage
will the voltmeter of Prob. 2.27 measure across one of the 10-kfJ resistors?
Repeat for two \-MQ resistors in series.
2.29 The D'Arsonval meter of Prob. 2.26 is used for the ohmmeter of Fig. 2.27.
What value of series resistance is required if E— 1.5V? What value of unknown
resistance will cause a one-quarter full-scale deflection?
2.30 Determine the color codes for resistors having the following resistance ranges:
(a)4.23-5.17 Q, (b) 6460-7140 ft, and (c) 3.135 -3.465 Mft.
i#
^
3
DEPENDENT SOURCES
The voltage and current sources of Chapters 1 and 2 are independent sources, as
defined earlier in Sec. 1.4. We may also have dependent sources, which are very
important in circuit theory, particularly in electronic circuits. In this chapter we shall
define dependent sources and consider an additional circuit element, the operational
amplifier, which may be used to obtain dependent sources.
We shall also analyze a few simple circuits containing resistors and sources, both
independent and dependent. As we shall see. the analysis is very similar to that
performed in Chapter 2. and the results may be used to construct a number of impor-
tant circuits, such as amplifiers and inverters, which will be defined in the chapter.
3.1 DEFINITIONS
A dependent or controlled voltage source is one whose terminal voltage depends on, or
is some other place in the circuit. A
controlled by. a voltage or a current existing at
voltage-controlled voltage source(VCVS) is a voltage source controlled by a voltage,
and a current-controlled voltage source (CCVS) .is one controlled by a current. The
symbol for a dependent voltage source with terminal voltage v is shown in Fig. 3.1(a).
A dependent, or controlled current source, symbolized by Fig. 3.1(b), is one whose
current is dependent on a voltage or a current existing elsewhere in the circuit. A
voltage-controlled current source (VCCS) is controlled by a voltage, and a current-
controlled current source (CCCS) is controlled by a current.
Figure 3.2 illustrates the four types of controlled sources and shows the voltage or
current on which they are dependent. The ju and ft are dimensionless con-
quantities
stants, commonly referred to as the voltage and current gain, respectively. The
constants /- and g have units of ohms and mhos, respectively.
49
50 Dependent Sources Chap. 3
'+N
Kt
6 6
(a) (b)
o-
+
V --
/iV, V = n,
(a) (b)
f J>
i = gv, f>..*l
(O (d)
FIGURE 3.2 (a) VCVS; (b) CCVS; (c) VCCS; (d) CCCS
Circuits containing dependent sources are analyzed in the same manner as those
without dependent sources. That is. Ohm's law
and Kirchhoff's voltage
for resistors
and current laws apply, as well as the concepts of equivalent resistance and voltage and
current division.
As an illustration of the procedure, let us find the current /in the circuit of Fig. 3.4.
Chap. 3 Dependent Sources 51
SV,
i
t
2n
Vva^-
- V, +
6
0.5i.
6V V,?6A
FIGURE 3.3 Circuit containing a dependent FIGURE 3.4 Dependent source example
source
-v, + 3v l
+ v2 = 6 (3.1)
f*i 3-V
and by Ohm's law we have
Vy = —2i, v2 = 6i (3.2)
2(-2z) + 6/ = 6
or / = 3 A. Thus the dependent source has complicated matters only to the extent of
requiring an extra equation, the first of (3.2).
As another example, let us find the voltage v in Fig. 3.5. Applying KirchhofTs
current law to the currents leaving the top node, we have
-4 + f, - 2/, + y = (3.3)
-4-* + *-0
otv= 12 V.
zn
-V^l
EXERCISES
3.2.1 In Fig. 3.3, let R =
{
3 Q, R2 = I Q, and /, = 2 A. Find i
g . Ans. 7 A
3.2.2 In Ex. 3.2.1, find the resistance seen by the source looking in terminals a-b.
Ans. § Q.
3.2.3 In Fig. 3.3, let /?, = R2 = r\j Q. Find the resistance seen by the source.
Ans. — j'-jj Q
A logical question at this point might be, How do we obtain dependent sources? One
answer is that they arise as parts of equivalent circuits of electronic devices operating
under certain conditions. Another answer is that they can be deliberately constructed
by means of certain electronic devices in conjunction with passive elements.
We shall not be interested here in undertaking a study of electronic devices.
However, there is one such device that is extremely useful in the construction of
dependent sources and whose ideal mathematical model is both simple and elegant.
This is the operational amplifier, or op amp, the ideal model of which we shall consider
in this section.
The symbol that we shall use for an operational amplifier is shown in Fig. 3.6. The
op amp is a multiterminal device, but for simplicity we shall show only the three
terminals indicated. Terminal 1 (marked — ) is the inverting input terminal, terminal 2
(marked +) is the noninverting input terminal, and terminal 3 is the output terminal.
The purposes of the terminals that are not shown include, in general, dc power supply
connections, frequency compensation terminals, and offset null terminals. We shall
not be interested in discussing these other terminals here, but the interested student
may find their purposes and how they are used in any op amp user's manual.
Operational amplifiers are commonly available in integrated circuit form and are
normally fabricated in packages having to4opamps.
8 to 14 terminals and containing 1
The operational amplifier has many characteristics that are important to designers,
but the ideal model of the op amp has only two properties that the circuit analyst needs
to know: that the currents into both input terminals are zero and that the voltage
between the input terminals is zero.
As an example of a an operational amplifier, let us consider Fig. 3.7.
circuit with
It is and the voltage v 3 considering v g to be a known
desired to find the current /'
,
generator voltage. Let us write KVL around the loop abca through the source. Since
the voltage across terminals a and b is zero, we have
v 1
~v = g (3.5)
Chap. 3 Dependent Sources 53
+ v2
2A 9 A
in
-o3
or i\ = vg . Applying KCL at node b and noting that the current into the negative
terminal of the op amp is zero, we have
(3.6)
1
or v 2 = —2v = — 2vg 1
. Next, KVL around loop cbdc through the 9-fi resistor yields
— v, + v2 + v3 =
or
v3 = v i
-v = 2 3v g (3.7)
i= v1 _3vt _v
9 " 9 "3a
Thus if vg = 12 cos 107, for example, we have i = 4 cos 10/. Also, we note that KCL
is not valid at the bottom node c. This is because node c (referred to as ground, to be
discussed in Chapter 4) is connected to the unshown power supply terminals.
insofar as vg i g v 3 and
, , , i are concerned, as shown in Fig. 3.8. Analysis of the equiva-
lent circuit yields exactly the same v 3 and i for the same v g and i g as in the case of Fig.
3.7. Thus the op amp has been used to obtain a controlled source with a gain of 3. (In
this case the controlled source is a VCVS, in which v g controls v 3 .)
ln=0
9n
Before we leave this section, let us observe that the op amp in Fig. 3.7 is being
operated in a feedback mode. That is, the output v 3 at node d is fed back to the
inverting input terminal through the 2-Q resistor. A practical op amp is a very high-
gain device and is generally never used without feedback. In cases when the feedback
is to one input terminal rather than to both, it is always to the inverting terminal, for
the simple reason that otherwise the op amp will not work. We are not interested here
in the reason for this, which is a consequence of the op amp's design. The interested
student, in all probability, will have occasion to thoroughly study op amps and their
construction in a later course in electronic devices.
Equations (3.6) and (3.7), which were the basis for the VCVS circuit of Fig. 3.8, are
independent of the current /in the 9-fi load of Fig. 3.7. This is true in general of VCVS
circuits of this type. To see this, let us consider the circuit of Fig. 3.9. The voltage v 2 is
the output voltage of the op amp and, as we shall see, is a function only of the input
voltage v l
and the two resistors.
Since there is no voltage across the input terminals of the op amp, we have
vba = v,. Also KVL around the loop abca containing v 2 yields
v hr = v,„ — v, = v, — v,
*!l_I_
v j
- v*
=
^1 ^2
Solving for v 2 , we have
v2 = nv x
( Mr
where
H=\
is therefore a VCVS with gain //. Since it is a three-terminal network
Figure 3.9
(the output and input share a common terminal) and the currents are zero into the
input leads of the op amp, an equivalent circuit may be drawn as in Fig. 3.10. We note
that since i?, and R 2 are nonnegative, we have fi 1. >
A special case of Fig. 3.9 is the case R 2 = (short circuit) and R = oo (open {
circuit), shown in Fig. 3.1 1. This circuit has = 1, or v 2 = v u and is called a voltage
fj,
follower; that is, v 2 follows v v It is also called a buffer amplifier because it may be used
+
H
'
+
+ o-
v,
«v,
Ri
3.9
to isolate, or buffer, one circuit from another. (The voltages at the two pairs of
terminals are the same, but no current can flow from one pair to the other.)
Let us consider next the circuit of Fig. 3.12. Applying KCL at the inverting input
terminal of the op amp, we have
Ri Ri
or
(3.10)
(Rec all tha t_the_¥o ltage acros input terminals of the op amp
are zero.)
^\AA^
vw -Q +
,—
v,
-o
—
This circuit is called an inverter because the polarity of v 1 is opposite that oiv x ,
as
seen in (3.10). It is also a VCVS, but in this case the input current i 1
is not zero, being
given by
(3.H)
*.
+ ^ +
-o +
R-,i
Z'l v2
We may also obtain dependent current sources from Fig. 3.12, which we redraw as
shown in Fig. 3.15. Since there is no current into the op amp terminals, we have
(3.12)
Ri
'i A AA^
+ o- vw
R2
Finally, substituting for/', in (3.12), we may redraw Fig. 3.16 as a VCCS, shown in
Fig. 3.17. In this case, g = 1/7?,.
v,
— o-
EXERCISES
3.4.1 If in Fig. 3.7 v g = 18 V, find v u v 2 and/. , Ans. 18 V, -36 V, 6 A
3.4.2 If in Fig. 3.12;, = 2 A, t- 2 = -10 V, and tf, = 2 ft, find R 2 and v x .
Ans. 5 ft, 4 V
3.4.3 If in Fig. 3.15 the terminals 3-4 are open, find i?, and ?;
43 if v t = \2V,i l
= 3 A,
and R2 = 2 ft. Ans. 4 ft, 6 V
PROBLEMS
3.1 Find the power delivered to the 8-ft resistor in the circuit shown.
10 V 5 V
PROBLEM 3.1
PROBLEM 3.2
+ V, -
PROBLEM 3.5
PROBLEM 3.7
3.8 Find i
g in the figure for Prob. 3.7 if R = 2 Q and / = 3 cos / A.
*--*&+#
(This circuit is called a summer, since the output voltage is the negative of a
weighted sum of the input voltages. Note that the result is independent of the
output connections at terminals a-b.)
Chap. 3 Dependent Sources 59
PROBLEM 3.9
3.10 In the figure for Prob. 3.9, connect a resistance of R = 5 kQ across terminals
a-b and find the resulting current iab (from a to b) if u, = 10 V, v 2 = 5 V,
^0=2 kfi, /?! = 1 kfi, and Rz = 2 kfi.
3.11 Tn the circuit given, show that regardless of the load at terminals c-d, we have
A\
' 1= ^' 2
Ri
oC +
PROBLEM 3.11
PROBLEM 3.13
60 Dependent Sources Chap. 3
3.14 Find R 2 /Ri in the figure for Prob. 3.13 so that i = 2 cos At mA.
3.15 Find i in the circuit shown if R = 6 Q and u g = 5 cos 10/ V.
R
-vw 3A< zni
PROBLEM 3.15
3.16 In the figure for Prob. 3.15, find v g so that R = 12 Q. and i = 100 mA.
**'
4
ANALYSIS METHODS
those that may be described completely by a single equation. The analysis of more
general circuits entails the solution of a set of simultaneous equations, as we shall
see in this chapter. As an example, the reader may have noticed that Prob. 2.23
required two equations in its solution. Also, most of the circuits of Chapter 3 generally
involved more than one equation, but the equations were of a type that were easily
solved.
In this chapter we ways of formulating and solving the
shall consider systematic
equations that arise in the analysis of more complicated circuits. We shall consider two
general methods, one based primarily on KirchhofT s current law and one on Kirch-
hoffs voltage law. As we shall see, KCL generally leads to equations in which the
unknowns are voltages, whereas KVL leads to equations in which the unknowns are
currents.
It should be evident from our work in previous chapters that a complete analysis
of a circuit can be performed by finding a relatively few key voltages and/or currents.
For example, in a simple circuit consisting of a single loop, a key variable is the
current, for if we know the current, we may find every voltage around the loop, and,
of course, the current around the loop is the current in every element.
In Sec. 4.1 we shall discuss the case where the selected unknowns are voltages.
Quite naturally, our choice of voltages should lead to a set of independent equations.
This technique will be referred to as nodal analysis. In Sec. 4.5 we shall consider mesh
analysis, in which the unknowns are currents.
In this chapter we shall discuss the techniques for selecting the voltages or currents
to befound and the formulation of the circuit equations. The justification of the
methods in the general case will be left until Chapter 6.
61
62 Analysis Methods Chap. 4
In this section we shall consider methods of circuit analysis in which voltages are the
unknowns to be found. A convenient choice of voltages for many networks is the set of
node voltages. Since a voltage is defined as existing between two nodes, it is convenient
to select one node in the network to be a reference node or datum node and then
associate a voltage or a potential with each of the other nodes. The voltage of each of
the nonreference nodes with respect to the reference node is defined to be a node
voltage. It is common practice to select polarities so that the node voltages are positive
relative to the reference node. For a circuit containing N nodes, there will be TV — 1
node some of which may be known, of course, if voltage sources are present.
voltages,
Frequently the reference node is chosen to be the node to which the largest number
of branches are connected. Many practical circuits are built on a metallic base or
chassis, and usually there are a number of elements connected to the chassis, which is
often then connected to the earth. The chassis may then be called ground, and it
becomes the logical choice for the reference node. For this reason, the reference node
is frequently referred to as ground. The reference node is thus at ground potential or
zero potential, and the other nodes may be considered to be at some potential above
zero.
Since the circuit unknowns are to be voltages, the describing equations are
obtained by applying KCL at the nodes. The currents in the elements are proportional
to the element voltages, which are themselves either a node voltage (if one element
node ground) or the difference of two node voltages. For example, in Fig. 4.1 the
is
reference node is node 3 with zero or ground potential. The symbol shown attached to
node 3 is the standard symbol for ground. The nonreference nodes 1 and 2 have node
voltages t>i and v2 . Thus the element voltage v 12 with the polarity shown is
v X2 = v x
-v 2
v 13 = v, -0 = v,
and
hz = v2 — = v2
These equations may be established by applying KVL around the loops (real or
imagined).
The application of KCL results in an equation relating node voltages. Clearly,
simplification in writing the resulting equations is possible when the reference node is
chosen to be a node with a large number of elements connected to it. As we shall see,
however, this is not the only criterion for selecting the reference node, but it is fre-
are those whose only sources are independent current sources. This is not always true,
as we we shall begin with an example of this type. In the network shown
shall see. but
in Fig. 4.2(a), there are three nodes, dashed and numbered as shown. [This is
easier to see in the redrawn version of Fig. 4.2(b).] Since there are four elements
connected to node 3, we select it as the reference node, identifying it by the ground
connection shown.
r)—Wv— R2
Q Q>
c z--T.-i^rzrz=:
(a)
Before writing the node equations, consider the element shown in Fig. 4.3, where
v x
and v 2 are node voltages. The element voltage v is given by
v = v,
= V\ -1>2
i
R R
or
i =G(v — l
v2)
where G= \jR is the conductance. That is, the current from node 1 to node 2 is the
difference of the node voltage at node I and the node voltage at node 2 divided by the
resistance R, or multiplied by the conductance G. This relation will allow us to rapidly
write the node equations by inspection directly in terms of the node voltages.
64 Analysis Methods Chap. 4
Now returning to the circuit of Fig. 4.2, the sum of the currents leaving node 1
'i + ii ~ i
g i
=
In terms of the node voltages, this equation becomes
G ]
vl +G 2 (v 1
- v2) - i gi =
We could have obtained this equation directly using the procedure of the previous
paragraph. Applying KCL at node 2 in a similar manner, we obtain
-i 2 + i3 + i g2 =
or
G2 (v 2 — ^i) + G> + 2 i
g2 =
Again, the G"s are the conductances (reciprocals of the 7?'s).
We could have equated the sum of currents leaving the node to the sum of currents
entering the node. Had we done so, the terms i gl and i gl would have appeared on the
right-hand side:
G l
v l
+ G (v - v =
2 l 2)
i gl
G 2
{v 2 -v +Gv = x ) 3 2 -i g2
+ G )v, - G v =
(G, 2 2 2 i gl (4.1)
-G v + (G + G )v =
2 x 2 3 2 -i g2 (4.2)
These equations exhibit a symmetry that may be used to write the equations in
the rearranged form by inspection. In (4.1) the coefficient of v is the sum of con- x
statement holds for (4.2) if numbers 1 and 2 are interchanged. Thus node 2 plays
the
the role in (4.2) of node 1 in (4.1). That is, it is the node at which KCL is applied. In
each equation the right-hand side is the current from a current source which enters
the corresponding node.
In general, in networks containing only conductances and current sources, KCL
applied at the kth node, with node voltage v k , may be written as follows. In the left
member the coefficient of v k is the sum of the conductances connected to node k, and
the coefficients of the other node voltages are the negatives of the conductances
between those nodes and node k. The right member of the equation consists of the net
current flowing into node k due to current sources.
To illustrate the process, consider Fig. 4.4, which is a portion of a circuit. The
dashed lines indicate connections to nodes other than node 2 (labeled v 2 ). At node 2 we
Chap. 4 Analysis Methods 65
1
F I GURE 4 . 3 Single element FIGURE 4.4 Part of a circuit
-<?!«, + (G, +G +G 2 3
)v 2 -G 2
v3 -G 3
i>
4
= i gl - i
g2 (4.3)
This may be checked by applying KCL in the usual way, equating currents leaving
node 2 to those entering node 2. The result is
G x
{v 2 - i\) +G 2 (v 2
- v3) +G 3
(v 2 - vt ) + /, 2 = / gi
4.2 AN EXAMPLE
To illustrate the method of nodal analysis, let us consider the circuit of Fig. 4.5. We
have taken the reference node as shown and labeled the nonreference nodes as v,, v 2 ,
and v 3 . We note that the conductances are specified for the resistors.
Since there are three nonreference nodes, there will be three nodal equations. At
node v u using the shortcut method, we have
4v, — V2 = 2 (4.4)
I
15
2U
^\r\r\j-
7A
© 313- 5A 3U IU 4U- 7A
node through the sources is 7 — 5 = 2. If we had used the conventional KCL method
at v i, we would have obtained
-v x + 6v — 2 2v 3 = 5
(4.5)
— 2v + 2 lv 3 = 17
We may solve (4.4) and (4.5) for the node voltages using any one of a variety of
methods for solving simultaneous equations. Two such methods are Cramer's rule,
which employs determinants, and Gaussian elimination. For the reader who is not
familiar with these two methods, a complete discussion is given in Appendices A
and B.
Using Cramer's rule, we first find the coefficient determinant, given by
4 -1
-1 6 145
-2
Then we have
2-1
5 6-2
v, = 17 -2 7 145
1 V
145 145
4 2
-1 5 -2
= 17 7 290
v, 2V
. 145 145
and
4 -1 2
-1 6 5
-2 17 435
= 3V
145 145
Now that we have the node voltages we may completely analyze the circuit. For
example, if we want the current i in the 2-15 element, it is given by
Equations (4.4) and (4.5) are symmetrical in a way that facilitates the writing of the
equations. This symmetry also is present in the general case. For example, the coeffi-
cient of v 2 in the first equation is the same as that of ?>, in the second equation. Also,
the coefficient of r, in the first equation is that off, in the third equation. Finally, the
coefficient of v 3 in thesecond equation is that of v 2 in the third equation. These
results follow from the fact that the conductance between nodes and 2 is that 1
between 2 and 1, the conductance between nodes and 3 is that between 3 and 1, etc. 1
This symmetry shows up also in the coefficient determinant A. Its diagonal ele-
ments. 4. 6. and 7. are the sums of the conductances connected to the three non-
reference nodes, and the off-diagonal elements are symmetrical about the diagonal.
The latter elements are, of course, the negatives of the conductances between nodes.
EXERCISES
4.2.1 Using nodal analysis, find v and v 2 in Fig. 4.2
x
(a) if R =
1
2CI, R2 = 2 Q,
R =4Q,
3 /„i = 2 A, and i <?2 -4 A. Ans. 7 V, 10 V
4.2.2 Repeat Ex. 4.2.1 with i g2 = 0. Check by using current division
Ans. 3 V, 2 V
4.2.3 Using nodal analysis, find v u v z and v } , . Ans. 8 V, -3 V, 5 V
EXERCISE 4.2.3
At first glance it may seem that the presence of voltage sources in a circuit complicates
the nodal analysis. We can no longer write the equations using the shortcut method
because we do not know the currents through the voltage sources. However, nodal
analysis is no more complicated and in many cases is even easier to apply when
voltage sources are present, as we shall see.
To illustrate the procedure, let us consider the circuit of Fig. 4.6. We have labeled
the nonreference nodes as v u v 2 v 3 and v, and have taken the sixth node as
,
Since there are five nonreference nodes, we need five equations. Without writing
any KCL equations we may note that we have, by inspection,
(4.6)
v< — Vd = V «2
We thus need only three KCL equations. To be systematic and at the same time elimi-
nate the need to know the currents in the voltage sources, let us enclose the voltage
sources by dashed lines as shown in Fig. 4.6. We may think of these surfaces as
generalized nodes or, as some authors prefer, super nodes. We recall that KCL holds
for such a generalized node as well as for an ordinary node. We have, therefore, two
generalized nodes and two regular nodes, labeled v 2 and v 3 a total of four nodes. ,
Thus we need only three KCL equations, which together with (4.6) constitute the
required set of five equations in the node voltages.
To complete the formulation of the nodal equations, let us apply KCL at nodes
v2 and v 3 and the generalized node containing v g2 . The first two are obtained as
before, resulting in
(C?, + G + G )v — G v — G v — G v =
2 4 2 x x 2 3 4 s
(4.7)
{G + G + G )v — G v - G i\ =
2 3 5 3 2 2 5
Finally, equating to zero the currents leaving the generalized node, we have
G4 (v -
s v2) +G 5
(v 4 - v3) + G 6 v5 = (4.8)
presence of the voltage source reduces the number of unknowns by one, we have
node at the top
labeled the left v
g
. the reference node being chosen as indicated. The
unknown node voltages are i\. v2 , and v 3 . Applying KCL at these nodes, we have
(G, -f G2 + G — G v — G v - G,v =
3 )i'i 2 2 3 3 g
-G 2 v + (G + G )v + P(v -v ) =
} 2 5 2 t 3
-G v + (G + G4 )v = fi{v, - v
3 t 3 3 3 )
These equations may now be solved for the unknown node voltages (v g is con-
sidered to be known). We note that the presence of the dependent source has destroyed
the symmetry that was present in the equations of the previous examples. This is true
in the general case involving dependent sources.
As a final example, let us consider Fig. 4.8, which contains independent and
dependent current and voltage sources. To simplify matters we have chosen the
reference node as shown and labeled the nonreference nodes, which in this case may
all be expressed in terms of only the two unknowns v x and i y (We have used Ohm's
_^L4_
+ Q _i_
2 ~ v* ~ (~ 5i y) - o
2 ^ '
and at the generalized node containing the dependent voltage source we have
^ = 4V, 1 A
We may now find all the currents and voltages in the circuit.
EXERCISES
4.3.1 Using nodal analysis, find yj if the element x is a 12-V independent voltage
source with the positive terminal at the top. Ans. —4 V
6V
2ft 3n ,Vo
6A 2A
EXERCISE 4.3.1
4.3.2 Repeat Ex. 4.3.1 if element x is a 0.5-A independent current source directed
downward. Ans. ~3V
4.3.3 Repeat Ex. 4.3.1 if element x is a dependent voltage source of 6 Wj V with the
positive terminal at the top. Ans. — V
1
Chap. 4 Analysis Methods 71
Nodal analysis very often is the best method of analysis when a circuit contains op
amps, because in electronic circuits the reference node is usually shown as grounded
and all other elements connected to the reference node are often shown individually
grounded. Thus the nodes are easily identified for the nodal method, but the loops are
not so easily visualized. Therefore a method based on loops, such as we shall consider
in the next section, is not so easy to apply. Also, quite often, only a relatively few nodal
equations arc required.
As an example, let us consider the VCVS of Fig. 3.9, redrawn as shown in Fig. 4.9.
The reference node is shown as grounded, so that the voltages v and v 2 of Fig. 3.9 are x
now node voltages, as shown. Since the voltage is zero between the input terminals of
the op amp and the currents are zero into the input terminals, we see that v 3 = «, and
that the node equation written at node v 3 is
i ,
t'i - v2
= (4.9)
R R,
As another example, let us consider the circuit of Fig. 4.10. We shall take the
ground to be the zero voltage reference point so that the inverting input terminal of
the op amp is also at zero potential, as labeled. We shall consider the voltage ?;, to be a
known input voltage and solve for the output voltage v 2 At node v 3 KCL yields .
,
(G, +G +G 2 3 -f G 4 )v — G
3 l
v 1
—G 4v2 =
using the shortcut procedure of Sec. 4.1. The sum of the currents entering the inverting
input node of the op amp is given by
G 3 v 3 + G s v2 =
72 Analysis Methods Chap. 4
(G, +G 2
G S
{G, G + G4 ) + G
3 3 G<
It is always fruitful to write nodal equations at the inverting input nodes of the op
amps, as we have done in this example. However, one generally avoids writing nodal
equations at output nodes of op amps because it is difficult to find the current out of
an op amp. There is no current into the input terminals, but because there are other
terminals not shown, the output terminal carries a current.
EXERCISES
4.4.1 Solve Ex. 3.4.1 using nodal analysis.
EXERCISE 4.4.3
Chap. 4 Analysis Methods 73
In the nodal analysis of the previous sections we applied KCL at the nonreference
nodes of the circuit. We shall now consider a method, known as mesh analysis, or loop
analysis, in which KVL is applied around certain closed paths in the circuit. As we
shall see. in this case the unknowns generally will be currents.
We shall restrict ourselves in this chapter to planar circuits, by which we mean
circuits that can be draw n on a plane surface in a way such that no element crosses any
other element. In this case, the planeis divided by the elements into distinct areas, in
the same way that the wooden or metal partitions in a window distinguish the window
panes. The closed boundary of each area is called a mesh of the circuit. Thus a mesh is
a special case of a loop, which we consider to be a closed path of elements in the
circuit passing through no node or element more than once. In other words, a mesh
is a loop that contains no elements within it.
As an example, the circuit of Fig. 4.11 is planar and contains three meshes,
identified by the arrows. Mesh contains the elements R R 2 R 3 and v gi mesh 2
1
x
, , , ;
In the case of nonplanar circuits (that is, those that are not planar), we cannot
define meshes. Thus in the analysis using KVL, the closed paths are loops. The
procedure is the same, of course, but the equations are not as easily formulated. We
shall consider this more general case in detail in Chapter 6.
/?,/, +R 3 I3 = v g\ (4.11)
R 2 I2 -R 3 I3 = -v g2 (4.12)
We define a mesh current as the current which flows around a mesh. The mesh
FIGURE4.11 Planar circuit with three meshes FIGURE 4.1 2 Circuit with two meshes
74 Analysis Methods Chap. 4
current may constitute the entire current in an element of the mesh, or it may be
only a portion of the element current. For example, in Fig. 4.12 the currents and i1
i2 are mesh currents, with the directions as shown. The element current is the mesh
current in R^ and R 2 , but the element current in R 3 is the composite of two mesh
currents.
In general, element currents are algebraic sums of mesh currents. This is illustrated
in Fig. 4.12 since the element current in R t
is
/i = h
that in R2 is
h= It —I = 2 /, - i2
iJj/j +R 3 (i 1
- i 2) = v gl
(4.13)
R2 i2 —R 3 (i l
— i2) = —v g2
These are the mesh equations of the circuit.
There is also a shortcut method of writing mesh equations which is similar to the
shortcut nodal method of Sec. 4.1. Rearranging (4.13) in the form
(/?, +R 3 )h - R = v el 3i2
-R 3 i, + (R 2 + R )i = -v g2
3 2
we note that in the first equation, corresponding to the first mesh, the coefficient of the
first current is the sum of the resistances in the first mesh, and the coefficients of any
other mesh current is the negative of the resistance common to that mesh and the first
mesh. The right member of the first equation is the algebraic sum of the voltage
sources driving the first mesh current in its assumed direction. Replacing the word
firstby the word second everywhere it appears in these last two sentences will describe
the second equation, and so forth. This shortcut procedure is a consequence of
selecting all the mesh currents in the same direction (clockwise in Fig. 4.12) and
writing KVL as the meshes are traversed in the directions of the currents. Of course,
the method applies only when no sources are present except independent voltage
sources.
As another example, let us return to Fig. 4.1 and define /,, i 2 and i 3 as the mesh 1 ,
currents shown in meshes 1, 2, and 3, respectively. Then applying the shortcut method
to mesh 1, we have
(R 1
+R + 2 i? 3 )/, -R 2i2
-R 3i3
= v gi
Chap. 4 Analysis Methods 75
i?,/, + R 2 (i - t
f 2) + /? 3 (/"i - h) = v aX
The analysis is completed by solving the three mesh equations for the mesh currents.
The same symmetry is present in the mesh equations as was noted in the nodal
equations. If Cramer's rule is to be used for solving for the currents, then the coeffi-
cient determinant to be calculated is
R + R2 + R
t 3
-/?, -R,
A = Rl + ^4 + ^5 — ^5
-R, —R 5 R +R + 3 s ^6
The diagonal elements are the sum of the resistances in the meshes, and the off-
diagonal elements are the negatives of the resistances common to the meshes corre-
sponding to the row and column of the determinant. That is, —R 2 in row 1, column 2
or in row
2, column 1 is the negative of the resistance common to meshes 1 and 2, etc.
Thus the determinant is symmetric about its diagonal.
EXERCISES
4.5.1 Using mesh analysis, find fj and i 2 in Fig. 4.12 if /?j = 1 Q, R2 = 2 Q,
R 3 = 4 Q, v gl = 21 V, and v g2 = 14 V. Ans. 5 A, 1 A
4.5.2 Repeat Ex. 4.5.1 with v g i = 0. Check by using voltage division.
Ans. 9 A, 6 A
4.5.3 Using mesh analysis, find /'[ and ;'
2 . Ans. 8 A, 5 A
EXERCISE 4.5.3
76 Analysis Methods Chap. 4
As in the case of nodal analysis of circuits with voltage sources, mesh analysis is
easier if there are current sources present. To illustrate this point, let us consider the
circuitof Fig. 4.13, which has two current sources and a voltage source.
With the mesh currents /,, i 2 and i 3 chosen as shown, it is clear that we need three
,
independent equations. Not all of these, however, have to be mesh equations. The
presence of the two current sources provides us with two constraints which we may
obtain by inspection:
h = —ie \
(4.14)
We need, therefore, only one more equation. Since it will have to come from
KVL, we need to select a closed path in which all the voltages are easily obtained.
That is, we need to avoid the current sources since their voltages are not readily
obtained.
If we imagine for a momenttwo current sources are removed, that is,
that the
opened, then we shall have two we already have two equations so
less meshes. But
there will still be enough meshes left for the required number of equations. Moreover,
the loops left (they may not be meshes) will have only resistors and voltage sources in
them, and therefore KVL is easily applied. We must stress that we are not taking the
current sources out. We are only imagining them out for a moment in order to locate
the loops around which KVL is to be applied.
Returning to Fig. 4.13 and imagining the current sources open for a moment, we
see that we have only one loop left, namely the loop containing v g3 /?,, R 2 and R 3 , , .
/?,(/,- i2) +R 2 (i 3
- i2) +R 3i 3
= v e3 (4.15)
The analysis of the circuit can now be completed by solving (4.14) and (4.15).
As a final example, let us apply loop analysis to the circuit of Fig. 4.8, which was
analyzed previously by the nodal method. The circuit is redrawn in Fig. 4.14. There
are four meshes and two constraints to be satisfied by the controlling variables v x and
i
y We shall need, therefore, six equations.
We may choose mesh currents and obtain their relations to the two current sources
as we did earlier in Fig. 4.13. However, to illustrate the use of loop currents and also
we have chosen as the unknowns currents/,, i 2 i 3
to simplify the resulting equations, , ,
FIGURE 4. 13 Circuit with two current sources FIGURE 4.14 More complex circuit
and one voltage source
/, =9
h + *2 +U= 1v x
i3 = i
y
From these results we may express all the loop currents in terms of v x and i
y . The
result is
U =9
(4.16)
/, = z,
,
_5v£ _o
current sources as open, we see the two loops to which we shall apply KVL. They are
abecdo. and bceb. The respective loop equations are
2/ 2 + 2 + 2/ 3 + 3/, + 1(i'i + Q=
and
l(/ 3 + /4 ) +2+ 2/ 3 =
78 Analysis Methods Chap. 4
3v
2
+ 5/,= -11
^4-3/ =7
The solution is given by
vx = A\, i=-lA
which checks with the result obtained earlier in Sec. 4.3. Clearly the nodal analysis
performed earlier was a simpler method.
The loop currents may now be found from (4.16), completing the analysis. Any
other current or voltage in the circuit may be readily obtained also.
EXERCISES
4.6.1 Using the methods of this section, find i u i2 and ; 3 , in Fig. 4.13 if R = t
1 Q,
R 2 = 3 Q, R 3 = 4 Q., i gl = 2 A, i gl = 8 A, and v g3 = 24 V.
Arts. -5 A, -2 A, 3 A
4.6.2 Repeat Ex. 4.6.1 if v g3 = 0. Ans. -8 A, -2 A,
4.7 DUALITY'
The reader may have noticed a similarity in certain pairs of network equations which
we have considered so far. For example, Ohm's law may be stated as
v = Ri (4.17)
or
i=Gv (4.18)
In the second case we have solved for i, of course, and used the definition G = \jR.
Another way of looking at these equations is to note that the second may be obtained
from the first by replacing v by /, i by v, and R by G. In like manner, the first may be
obtained from the second by replacing i by v, v by /, and G by R.
Similarly, in the case of series resistances R R2 t
, , . . . , R n,
the equivalent resistance
was shown in Sec. 2.3 to be
R =
S
R, +R + 2 ... +R n (4.19)
Chap. 4 Analysis Methods 79
Gp = G, +G + 2 . . . +G n (4.20)
There is thus a definite duality between resistance and conductance, current and
voltage, and series and parallel. We acknowledge this by defining these quantities as
duals of each other. That is. R is the dual of G, i is the dual of v, series is the dual of
parallel, and vice versa in each case.
Another simple case of dual equations is
v = (4.21)
i = (4.22)
In the general case, an element described by (4.21) is a short circuit, and one described
by (4.22) is an open circuit. Thus short circuits and open circuits are duals. There are
also other dual quantities, as we shall see later.
Every equation in circuit theory has a dual, obtained by replacing each quantity in
the equation by its dual quantity. If one equation describes a planar circuit, then the
other equation describes the dual of the circuit, or the dual circuit. (As the reader may
see in a later course, nonplanar circuits do not have duals.) For example, consider the
circuit of Fig. 4.15. The mesh equations are given by
+R -R
(/?, 2 )i l 2i 2
= vg
(4.23)
-R + (R + 7?
2 i, 2 3 )/ 2
=
To obtain the dual of (4.23), we simply replace the R's by G's, the fs by v's, and v
by /'.
The result is
((7, + G )v, - G v =
2 2 2 i
g
(4.24)
— G2v + {G + G )v =
t 2 3 2
These are the nodal equations of a circuit having two nonreference node voltages, v x
and v z three conductances, and an independent current source i g From our shortcut
, .
procedure for nodal equations we see that G^ and G 2 are connected to the first node,
G2 is common to the two nodes, G 2 and G 3 are connected to the second node, and i
g
enters the first node. Since (7, and i g are not connected to the second node and G 3 is
not connected to the first node, these elements are connected to the reference node.
Such a circuit is shown in Fig. 4.16 and is a dual of Fig. 4.15.
Figure 4.15 may be described as vg in series with R l
and the parallel combination
of R 2 and R 3 . Replacing the quantities in this statement by their duals, we see correctly
80 Analysis Methods Chap. 4
In this example we note that nodes are duals of meshes and vice versa. This is true
in general because of the duality between the mesh currents and the nonreference node
voltages. The reference node is the dual of the boundary of the region outside the
circuit, or what we might call the outer mesh.
This last duality (between nodes and meshes) gives us a procedure for finding a
dual of a given circuit. We may place one node of the dual network to be obtained
inside each mesh, thereby assuring us of the correspondence between each nonreference
node and a mesh. Then we may place the reference node outside the circuit (corre-
sponding to the outer mesh). Connecting these nodes together by elements drawn
through the elements of the original circuit assures us of correspondence between
elements. If these last elements drawn are the duals of those they cross, then the
circuit obtained is a dual of the original circuit.
As an example, the circuit of Fig. 4.15 is redrawn using solid lines in Fig. 4.17. The
dashed circuit, superimposed in accordance with the dual circuit procedure, is its dual.
It should be clear that the latter circuit is the same as that of Fig. 4.16.
In the case of independent or dependent sources, the polarity of the dual source is
specified when the dual circuit is obtained from the dual equations, as was the case in
Fig. 4.16. However, in the geometric method of Fig. 4.17, the polarity of i g cannot be
determined since the mesh current directions are not shown and thus are completely
arbitrary.
~<9 ^ / u3
If the mesh currents or node voltages are shown in the original circuit, then the
polarity of a source obtained in the dual circuit may be determined from the concept
of driving a node or a mesh. We shall say that a mesh is driven by a source in it if the
polarity of the source is such as to drive the current in the direction of the mesh
current. A node is driven by a source connected to it if the polarity of the source is
such as to apply voltage to the node, i.e., send current toward the node. Thus the dual
source will drive or not drive a mesh or a node in accordance with whether the source
it crosses in the original circuit drives or does not drive the corresponding node or
mesh. This is consistent, as it must be, with the two sets of dual equations of the
circuits.
In general, when we have analyzed one circuit, the numerical values of its voltages
and currents are the same as those of their duals in the dual circuit. When we solve
one circuit, therefore, we have really solved two. As an illustration, the numerical
values of/', and i 2 in Fig. 4.15 are, respectively, the same as those of f, and v 2 in
Fig. 4.16.
As a final note in this chapter, we shall not attempt to give a dual of an op amp.
We may. of course, obtain duals of circuits containing dependent sources which are
equivalent to circuits with op amps.
EXERCISE
4.7.1 Find the dual of the circuit in the figure for Ex. 4.5.3.
PROBLEMS
4.1 Using nodal analysis, find the current / in the circuit shown.
4A
PROBLEM 4.1
—
4.2 Using nodal analysis, find the voltage v in the circuit shown,
2A
-0-
in 2ft
-wv- ^\^
PROBLEM 4.2
4.3 Repeat Prob. 4.2 if the 6-A source is replaced by an independent voltage source
having a terminal voltage of 10 V with the positive terminal at the top.
2V
3ft
-0-
PROBLEM 4.4
12V 3mA
PROBLEM 4.5
Chap. 4 Analysis Methods 83
2 mA
PROBLEM 4.6
18V
2a
PROBLEM 4.7
4.8 Using nodal analysis, find the power delivered to the 4-Q resistor in the circuit
shown.
PROBLEM 4.8
84 Analysis Methods Chap. 4
PROBLEM 4.9
12V
6mA 2kn
CD i
mA
PROBLEM 4.10
4.16 Using mesh or loop analysis, find v in the circuit shown. The sources x, y, and z
are independent current sources with arrows pointing upward and currents of
10, 15, and 5 A, respectively.
3n
20, 4ft
-A/W-
3fi
-V\A/-
PROBLEM 4.16
Chap. 4 Analysis Methods 85
4.17 Solve Prob. 4.16 if source z is changed to a 12-V voltage source with the positive
terminal at the top.
4.18 Solve Prob. 4.16 using any method if and z are independent voltage sources
x, y,
w nil the positi\e terminals at the top and terminal voltages of 39, 10, and 20 V,
respectively.
-0-
2mA
|3V
3KQ
|4V
I Kit
PROBLEM 4.19
4kU
-AA/V—
3mA
25V
4kQ
5mA
© •
3 k n
2kn
6 IOmA
PROBLEM 4.20
12V
PROBLEM 4.21
86 Analysis Methods Chap. 4
4.22 Find v t
in the circuit shown.
PROBLEM 4.22
»2 %-5x I0*lg
PROBLEM 4.23
4.24 Find v in terms of the node voltages v u v2 and v 3 and the resistances in the
,
circuit shown. (This circuit is a summer, like that of Prob. 3.9, with an additional
voltage source.)
R
"I o-
n2
v 0-
2
-o v o
n3
"3 0-
PROBLEM 4.24
Chap. 4 Analysis Methods 87
IkO
-AAA—
2kf! Ikfi
v
i
o aaa^- AAA f
-OVg
I k
4 2kn
5 Ikn
PROBLEM 4.25
G„
Aaa-
"3
AAA^
G2
I O » AAA i
-O u 2
-Wfv-
G,
PROBLEM 4.26
4.27 Find the input resistance R ia , in terms of the other resistances, in the circuit
shown.
PROBLEM 4.27
'4.28 Draw the dual of the circuit of Prob. 4.1, and find v in the dual circuit corre-
sponding to the unknown i in the original circuit.
88 Analysis Methods Chap. 4
*4.29 Draw a circuit and its dual if the mesh equations of the circuit are given by
- 2/ 2 = 4
10/,
-2/, + 8/ 2 - f = 3
-iz + ll/ = -6 3
*4.30 Draw the dual of the circuit of Prob. 4.16. Identify i in the dual circuit corre-
sponding to v in the original circuit and show that they have the same numerical
values.
5
NETWORK THEOREMS
by the use of certain network theorems. For example, if we are interested only in what
happens to one particular element in a circuit, it may be possible by means of a
network theorem to replace the rest of the circuit by an equivalent and simpler
circuit.
v = Ri
and we considered circuits that were made up of linear resistors and independent
sources. We defined dependent sources in Chapter 3 and analyzed circuits containing
both independent and dependent sources in Chapter 4. The dependent sources that we
89
90 Network Theorems Chap. 5
y = kx (5.1)
where A: is a constant and the variables x and y were either voltages or currents. Clearly,
Ohm's law is a special case of (5.1).
In (5.1) the variable y is proportional to the variable x, and the graph of >> versus x
is a straight line passing through the origin. For this reason, some authors refer to
elements which are characterized by (5.1) as linear elements.
For our purposes we shall define a linear element in a more general way, which
includes (5.1) as a special case. If x and y are variables, such as voltages and currents,
associated with a two-terminal element, then we shall say that the element is linear if
multiplying x by a constant K results in the multiplication of y by the same constant
K. This feature is called the proportionality property and evidently holds for (5.1) since
Ky = k(Kx)
Thus not only is an element described by (5.1) linear, but, in addition, elements
described by relations of the forms
$-«, y = »% (5-2)
The describing equations of a linear circuit are obtained by applying KVL and
KCL, and therefore they contain sums of multiples of voltages or currents. For
example, a loop equation is of the form
a v
x x
+ a2v2 + . . . + a n vn =f (5.3)
where/is the algebraic sum of the voltages of the independent sources in the loop, the
v's are the voltages of the remaining loop elements, and the o's are or ±1. To
illustrate, KVL around the loop shown in the circuit of Fig. 5.1 yields
v l
+ v2 - v3 = v gl - v g2
In this case, a x
= a2 = 1; a3 = —I; the other a's, if any, are all zero; and
Chap. 5 Network Theorems 91
v x
= 2ij, v2 = 5/ 2 , v3 = 3/ 6
Since by KCL the current to the right in the 2-Q resistor is /' — /„,,
l
«2' KVL around the
left mesh yields
2(/ - i
t2 ) + 4/ = v*i (5.4)
from which
I
—
= l£l-
^
I '_£2
(5.5)
3
If v gl = 18 V and i
g2 = 3 A, then / =3+ = 1 4 A. If we double v gl to 36 V and / Kl
to 6 A, then i is doubled to 8 A.
2fl
© 4a
© g2
FIG U R E 5.1 Loop of a linear circuit FIGURE 5.2 Linear circuit with two sources
1 V
92 Network Theorems Chap. 5
and see where it leads us. Referring to the figure, this assumption on v gives
x
/, =2A
and i2 = 1 A, and therefore
it = i, 3A
Proceeding down the ladder toward the source, we have, by KVL and KCL,
3/, 9V
V3 = V \ +V = 2 10 V
2A
5
is = i3 + /'
4 =5A
v4 = l(/ 5 ) = 5V
and, finally, if the guess that v x = V 1 is correct,
v. = v-. v* = 15 V
Our guess was not correct, since actually v g = 45 V, but in view of the laws of
probability this should not surprise us. However, by the proportionality relation, if a
15-V source gives an output v x
= 1 V, then our 45- V source will give three times as
much, so that the correct answer is
v, =3V
This method of assuming an answer for the output, working backwards to obtain
the corresponding input, and adjusting the assumed output to be consistent with the
actual input, by means of the proportionality relation, is particularly easy to apply to
the ladder network. There are other circuits to which it also applies, but the ladder
network is one of the most-often-encountered circuits, and the method is worthwhile
for it alone.
A nonlinear circuit is, of course, one that is not linear. That is, it has at least one
element whose terminal relation is not of the form of (5.1) or (5.2). An example is
given in Ex. 5.1.4, for which it is seen that the proportionality property does not
apply.
/
EXERCISES
5.1.1 Show that in Ex. 4.5.3 if all the source voltages are doubled, then the loop
currents are doubled.
5.1.2 Show that in Ex. 4.4.3 if the source voltage is doubled, then the response / is
doubled.
5 in Fig. 5.3.
Am. 6 A, 27 V, 9 A, 30 V, 15 V, 15 A
5.1.4 A circuit is made up of a voltage source v g a , 2-fi linear resistor, and a non-
linear resistor in series. The nonlinear resistor is described by
v = i
2
where v is the voltage across the resistor and ;', which is constrained to be non-
negative, is the current flowing into the positive terminal. Find the current flow-
ing out of the positive terminal of the source if (a) »g = 8V and (b) v g = 16 V.
Note that the proportionality property does not apply.
Am. (a) 2 A, (b) 3.123 A
5.2 SUPERPOSITION
In this section we shall consider linear circuits with more than one input. The linearity
property makes it possible, as we shall see, to obtain the responses in these circuits by
analyzing only single-input circuits.
To see how this may be accomplished, let us first consider the circuit of Fig. 5.2,
which was analyzed in the previous section. The output i satisfied the circuit equation
(5.4), which we repeat as
/ = 5jl + if (5.7)
also given earlier, we see that i is made up of two components, one due to each input.
If/, is the component of /'due to v gX alone, i.e., with i gl = 0, then by (5.6) we have
Similarly, if 2 is the component of/ due to i gl alone, i.e., with v gl = 0, then by (5.6)
we have
2(/ 2 - i gl ) + 4/ 2 = (5.9)
94 Network Theorems Chap. 5
i = h + i%
— v g\
t tol
Alternatively, we may obtain the components of/ directly from the circuit of Fig.
5.2. To find i we need to make the current source i gl zero. Since i g2 =
l
is the equa-
tion of an open circuit, this is accomplished by replacing the current source in the
circuit by an open circuit. This operation of making a source zero is sometimes
referred to, rather grimly, as "killing" the source, or making the source "dead." The
resulting circuit, in this case, shown in Fig. 5.4(a), from which it is easily seen, using
is
Ohm's law, that /,, the component of i due to v gl alone, is given by the first equation
of (5.11).
(al (b)
FIGURE 5.4 Circuit of Fig. 5.2 with (a) the current source dead and
(b) the voltage source dead
To find i2 , the component of/ due to i g2 alone, we must have v gl = 0, the equation
of a short circuit. Thus to kill a voltage source such as v gl we replace it by a short
,
circuit, as shown in Fig. 5.4(b). Using current division it is easy to see that i 2 is given by
the second equation of (5. 11).
The method we have illustrated with this example is called superposition, because
we have superposed, or algebraically added, the components due to each independent
source acting alone to obtain the total response. The principle of superposition
Chap. 5 Network Theorems 95
applies to any linear circuit with two or more sources, because the circuit equations are
linear equations (first-degree equations in the unknowns). This may be seen from (5.3)
and the nature of the v-i relations for the linear elements. In particular, in the case of
resistive circuits, the response may be found using Cramer's rule and is clearly a sum of
components, one due to each independent source alone. (Expanding the numerator
determinant by cofactors of the column containing the sources clearly reveals this.)
The linearity of the circuit equations enabled us to add equations (5.8) and (5.9)
and to see in (5.10) that the response was the sum of the individual responses. For
example, we are able to say that
As a second example, let us find the voltage v in the circuit with three independent
sources,shown in Fig. 5.5. To illustrate the earlier statement that the response is a sum
6V
en
of components, each due solely to an independent source, we shall first solve the
circuit by conventional means. Then to contrast the methods we shall use superposi-
tion.
To illustrate the role each source plays in the response we shall label the sources as
J
g\ 6V
'*2 (5.12)
Hi
= 18V
Assigning the mesh currents as indicated, the mesh equations are then given by
8/'i
— 2i 2 = —v *3
(5.13)
2/j + 5/ 2 = 3/ *:
8 - v *i '
h = -2 v gi - 3/, 2 2
v *i
2 . 1
V* 3
8 -2 -
9
"
T8
-2 5
so that
v = Ki - ',2 + K3
Thus we see that v, as well as i2 , is a sum of components due to the individual sources.
Finally, substituting (5.12) into the expression for v, we have
w=4-2+3=5V (5.15)
v = *y, + v2 + v3 (5.16)
where ?',component due to the 6-V source alone (the 2- A and 18-V sources
is the
dead), v 2 component due to the 2 -A source alone (the two voltage sources dead),
is the
and i>is the component due to the 18-V source alone (the 6-V and 2- A sources dead).
3
The circuits showing v v 2 and v 3 are given in Figs. 5.6(a), (b), and (c), respectively.
r
, ,
In Fig. 5.6(a), killing the 18-V source ties nodes b and d together; in Fig. 5.6(b),
killing the two voltages ties nodes a and c together and b and d together; etc.
Chap. 5 Network Theorems 97
2A
e
a,c
en V^/\J\|
en
(c)
v, =4V
r2 = -2V
6 A
P =
which is a quadratic, and not a linear, expression. Therefore superposition will not
apply to obtaining power directly. That is, we cannot find the power due to each
independent source acting alone and add the results to obtain the total power. How-
ever, we can find v by superposition and subsequently obtain p.
Letting v x
be the component of v when the 12-V source is acting alone [Fig. 5.8(a)]
and v 2 be the component of v due to the 6-A source alone [Fig. 5.8(b)], we have, as
before,
v = v, + v2
We note that the current i driving the dependent source also has components, which
we have labeled /, and i2 , due to each source alone. Solving for v in Fig. 5.8(a) and v 2 x
= (15)^ = 75 w
'i 12 V '2
o—,
(o) (b)
FIGURE 5.8 Circuit of Fig. 5.7 with (a) the current source killed and
(b) the voltage source killed
There are at least three things that are illustrated by the last example. First, as
meshes in Fig. 5.7. the original circuit, and one mesh current is known. Therefore only
one application of KVL is needed, around the left mesh, to solve for the current in the
1-fi resistor. This is also exactly the case in the circuit of Fig. 5.8(b).
EXERCISES
5.2.1 Solve Ex. 4.2.3 using superposition.
5.2.3 Find the power delivered to the 3-fi resistors in Figs. 5.8 (a) and (b), and show
that their sum is not equal to the total power delivered to the 3-fi resistor in
Fig. 5.7. Am. 12 W, 27 W
In the previous section .of this chapter we saw that the analysis of some circuits could
be greatly simplified by applying the principle of superposition. However, as we saw
in the last example of Sec. 5.2, superposition alone may not reduce the complexity of
the problems, but its use may lead to additional work. In this section we shall consider
Thevenins and Norton's theorems, which will in many cases be applicable and greatly
simplify the circuit to be analyzed.
We shall assume that the circuit to be considered can be separated into two parts,
as shown in Fig. 5.9. The part denoted as circuit A is a linear circuit containing
resistors, dependent sources, or independent sources. Circuit B may contain nonlinear
elements. We also add the constraint that any dependent source in either circuit must
have its controlling element in the same circuit. That is, no dependent source in
circuit A can be controlled by a voltage or current associated with an element in
circuit B and vice versa. The reason for this will become clear in the development which
follows.
We shall show that we can replace circuit A by an equivalent circuit containing a
source and a resistor in such a way that the voltage-current relations at terminals a-b
remain the same. Since our aim is to maintain the same terminal relations at a-b,
clearly, from the standpoint of circuit A, we may obtain the same effect if we replace
circuit B by a voltage source of v volts with the proper polarity, as shown in Fig. 5.10.
Insofar as circuit A is concerned, it has the same terminal voltage, and since circuit A
itself is unchanged, the same terminal current must flow. We have now obtained a
linear circuit and can consequently make use of all the properties we have established
for such circuits.
100 Network Theorems Chap. 5
Linear + Linear
Circuit
Circuit V Circuit
A B
A
b
i = ix + 4 (5.17)
where i x
is produced by the voltage source v with the network A dead (all its indepen-
dent sources killed) and iK is the short-circuit current produced by any sources inside
circuit A with v killed (replaced by a short circuit). These two cases are shown in Figs.
5.11(a) and (b).
__
Circuit Circuit
dead
A
A A
alive
R th
(a) (b)
Since the independent sources are dead in circuit A [Fig. 5.11(a)], from the termi-
nals of the source v we see only a resistive circuit, the equivalent resistance of which we
shall call R tb . By Ohm's law, we therefore have
v
(5.18)
1 = ~ir + ls (5.19)
0- -#* + *.
or
(5.20)
v = -R J + t
v, (5.21)
The relations given in (5.19) and (5.21) may be used to obtain two very useful
circuitswhich are equivalent to circuit A. Historically, the first of these was Thevenin s
equivalent circuit, named in honor of the French telegraph engineer Charles Leon
Thevenin (1857-1926), who published his results in 1883. [The circuit might have been
more appropriately named for the great German physicist H. S. F. Helmholtz (1821-
1894), who gave a restricted case of it in 1853.]
The Thevenin equivalent circuit is simply one which is described by (5.21) with
terminal voltage v and terminal current /, oriented as in Fig. 5.9. To draw the circuit
we note that v is the sum of two terms, which therefore must represent two elements in
series whose terminal voltages add up The first term evidently corresponds to a
to v.
resistance i? th called the Thevenin resistance, and the second term corresponds to a
,
voltage source with terminal voltage oc The result is shown in Fig. 5.12, where the
^' .
dashed lines represent the connections to the external circuit B of Fig. 5.9. Analysis of
the circuit will show that (5.21) is satisfied. The statement that Fig. 5.12 is equivalent
at the terminals to Fig. 5.9 is known as Thevenin s theorem.
Another circuit that is equivalent to circuit A
of Fig. 5.9 is obtained from (5.19).
This circuit is the duaFof the Thevenin circuit
and is called the Norton equivalent
circuit in honor of the American engineer E. L. Norton (1898- ), whose work was
published some 50 years after Thevenin's. From (5.19) we see that is the sum of two /'
terms, which must then represent two parallel elements whose currents add up to i. The
first term evidently arises from the Thevenin resistance R lh and the second term ,
corresponds to a current source /sc The result is shown in Fig. 5.13, and the statement
.
"th
6
FIGURE 5.12 Thevenin equivalent FIGURE 5.13 Norton equivalent circuit of
circuit of Circuit A of Fig. 5.9 Fig. 5.9
102 Network Theorems Chap. 5
of equivalence of Figs. 5.9 and 5.13 is Norton's theorem. Again, the dashed lines
represent the connections to the external circuit B.
As an example, let us find the Thevenin and Norton equivalent circuits for the
network to the left of terminals a-b in Fig. 5.14. Then using the results, let us obtain
the current i, as shown, in terms of the load resistance R.
6 V
4-Q-
FIGURE 5.14 Circuit with a variable load resistor
obtain the Thevenin circuit we need to find R th and vx The Thevenin resistance
To .
R lh found from the dead circuit (the two independent sources made zero), shown in
is
th
(o)
FIGURE 5.15 Circuits for obtaining the Thevenin circuit of Fig. 5.14
The open-circuit voltage v x is obtained from Fig. 5.15(b). Since the terminals a-b
are open, the voltage v oc is across the 3-Q resistor. Labeling the nodes as shown, with
node b as reference, and writing a nodal equation at the generalized node, shown
dashed, we have
"
3
or
^-tf = 6V J
The Thevenin equivalent circuit, with the load R connected, is shown in Fig. 5.16.
(a
t
Chap. 5 Network Theorems 103
We note that the polarity for i ^— 6 V is such that the correct voltage polarity
results at terminals a-b when they are opened.
The current i in Fig. 5. 14 is the same as that of Fig. 5.16, which in the latter case is
readily seen to be
1
R 1 4
We may use this result to find the load current for any load R that we choose.
To obtain the Norton equivalent circuit we use R lh = 4 Q, as before, and calculate
/sc . We may short terminals a and b and calculate isc from the resulting circuit, or we
may use the v x we already have and get /'
sc
from (5.20). In the latter case we have
The Norton equivalent circuit, with the load R connected, is shown in Fig. 5.17.
Using current division, we have, as before,
C M C
FIGURE 5.16 Loaded Thevenin equi- FIGURE 5.17 Loaded Norton equivalent of
valent of the circuit of Fig. 5. 14 the circuit of Fig. 5.14
The direction of the 1.5-A source is such that when R is replaced by a short,
/ sc = 1.5 A has the correct direction. In this case and in the Thevenin case of Fig. 5.16,
it is a simple matter to place the sources correctly, but in a complex example some care
needs to be exercised so that the polarities are correct.
Let us now consider an example containing a dependent source, such as the circuit
of Fig. 5.18(a). Suppose we want the Norton equivalent circuit at terminals a-b. We
shall need, in this case, RA defined for the dead circuit of Fig. 5.18(b) and / sc shown in
Fig. 5.18(c).
We may find /'
10 h — i*
2i, V
i <r +> T
- ;
— ^AA/
V °
-r—^VW
3n
—
4a 6n
(a) <th
(b)
10 A
(c)
FIGURE 5.1 8 fa,) Circuit to be analyzed, with (b) its source killed and
(c) its terminals short-circuited
Eliminating i u we have
L= 5 A (5.22)
^oc = 6/!
As a final example, let us find the Thevenin equivalent of the circuit of Fig. 5.20.
To begin with, we see by inspection that since there is no independent source present,
we must have
Voc =L= (5.23)
Also, the dead circuit is the given circuit itself, so that R th is simply the resistance seen
at the terminals of Fig. 5.20. In view of (5.23), we cannot use v oc = i? th / sc to get i? th , as
we did in the previous example. The only recourse we have is to excite the circuit at its
2i, V
v-2i,
4n 6n> , (D
IA
4fl
'th
FIGURE 5.20 Circuit with a dependent FIGURE 5.21 Circuit of Fig. 5.20 excited by a
source current source
For example, suppose we excite the circuit with a 1-A current source, as shown in
Fig. 5.21. Then we have
v
R t
where v is the resulting terminal voltage. Taking the bottom node as reference, the
nonreference node voltages are as shown. A nodal analysis yields
v — 2/,
= 1
6
where
3n
EXERCISES
5.3.1 Replace the circuit to the left of terminals a-b by its Thevenin equivalent and
find i. Ans. v x = 24 V, R th = 6 Q, i =3A
24V
EXERCISE 5.3.1
5.3.2 Replace everything in the figure for Ex. 5.3.1 except the 3-£2 resistor by its
In Chapter 1 we defined independent sources and pointed out that they were ideal
elements. An ideal 12-V battery, for example, supplies 12 V between its terminals
regardless of the load connected to the terminals. However, a real, or practical, 12-V
battery supplies 12 V when its terminals are open-circuited and supplies less than 12 V
Chap. 5 Network Theorems 107
as current is drawn through the terminals. A practical voltage source thus appears to
have an internal drop in voltage when current flows through its terminals, and this
internal drop diminishes the voltage at the terminals.
We may represent a practical voltage source by the mathematical model of Fig.
5.23, consisting of an ideal source v g in series with an internal resistance R g The
.
voltage v seen at the terminals of the source now depends on the current i drawn from
the source. The relationship is easily seen to be
RJ (5.24)
A PRACTICAL
VOLTAGE
SOURCE
Thus under open-circuit conditions (/' = 0), we have v = vg , and under short-circuit
conditions (v = 0), we have i = vjR g
. If Rg > 0, as it is for a practical source, the
source can never deliver an infinite current, as an ideal source can.
For a given practical voltage source (fixed values of v g and Rg in Fig. 5.23), the
load resistance RL determines the current drawn from the terminals. For example, in
Fig. 5.23 the load current is
/ = (5.25)
Rg + R L
DEAL SOURCE
v — R,v
L"g
(5.26)
R g + Rl
Therefore as we vary RL both i and v vary. A sketch of v vs. RL is shown in Fig. 5.24,
along with the ideal case, which is dashed. For large values of RL relative to Rg , v is
very nearly equal to the ideal value of v g (If . RL is infinite, corresponding to an open
circuit, then v is vg .)
We may replace the practical voltage source of Fig. 5.23 by a practical current
source by means of Norton's theorem. The result is shown in Fig. 5.25 and is seen to
be an ideal current source in parallel with the internal resistance. The two sources in
Figs. 5.23 and 5.25 are equivalent at the terminals, of course, if Rg is the same in both
cases, and
vg =R i
g g (5.27)
RJ g (5.28)
R. + Rr
A PRACTICAL
CURRENT
SOURCE
Fig. 5.27. We could solve the problem in a number of ways, such as replacing every-
thing except the 4-Q resistor by its Thevenin equivalent and then finding i. However,
we shall illustrate instead the method of successive transformation of sources.
Let us begin by replacing the 32-V source and internal 3-Q resistance by a practical
current source of a 3-fi internal resistance and a ^-A ideal source. Then let us replace
the 4-A source and the internal 2-Q resistance by a voltage source of 2(4) = 8 V and a
Chap. 5 Network Theorems 109
IDEAL SOURCE
3A 4a in
32 V
Q 6A 6A 2H
© 4A
32
We may now combine and 6-Q resistances and the series l- and 2-Q
the parallel 3-
resistances, as shown and repeat the source transformation procedure.
in Fig. 5.29(a),
Continuing the process, as shown in Figs. 5.29(b), (c), and (d), we finally arrive at an
equivalent circuit (insofar as is concerned) which can be analyzed by inspection. In
/'
64 16
/ = 3 J- = 2A
2 + 4 + 2
This procedure may seem unduly long, but it should be observed that most of the steps
may be carried out mentally.
— -
(o)
¥v©
(b)
i
/
2A
WVL —
i-». 4A
j vw-
T*© 2a
on*
(c)
2n 2X1
i
—VW^
^V
© © S*
(d)
obtain equivalent sources. For example, ifwe are interested only in i in Fig. 5.29(d), we
may combine the three series resistors, as we know, but we may also combine the
series sources. They represent a net source of ^—^= 16 V with a polarity like that
Chap. 5 Network Theorems 111
of the larger source. Thus an equivalent circuit insofar as i is concerned is that of Fig.
5.30. Similarly, we may combine parallel current sources to obtain an equivalent
source.
EXERCISES
5.4.1 By source transformations, replace the network to the left of terminals a-b in
the figure for Ex. 5.3.1 by an equivalent network having only one resistor R.
From this network, find /'
as shown. Am. R = 6 Q, / = 3 A
5.4.2 Convert all the sources in the figure for Ex. 4.2.3 to voltage sources and find u 3 .
Arts. 5 V
5.4.3 In the figure for Prob. 4.5, convert the network to the left of terminals a-b to a
single practical current source by source transformations and find v. Ans. 10 V
There are many applications in circuit theory where it is desirable to obtain the
maximum possible power that a given practical source can deliver. It is very easy, using
Thevenin's theorem, to see what maximum power a source is capable of delivering and
how to load the source so as to obtain this maximum power. This will be the subject of
this section.
Let us begin with the practical voltage source shown earlier in Fig. 5.23 with a load
resistance R L The
. power p L delivered to the resistor RL is given by
tion of RL . To maximize p L we can make dpJdR L = and solve for R L From (5.29) .
112 Network Theorems Chap. 5
we obtain
+ R L - 2(R g + R L )R,
(R g )
2
dR L (R g + RlY
(R g - R L )vt = Q
(5.30)
(R g + RLy
which results in
RL = R g (5.31)
d lp L
dR\ SRI
and therefore (5.31) is the condition which maximizes p L We see, therefore, that the .
maximum power is delivered by a given practical source when the load R L is equal to
the internal resistance of the source. This statement is sometimes called the maximum
power transfer theorem. We have developed it for a voltage source, but in view of
Norton's theorem it also holds for a practical current source.
The m aximum po wer th at the practical voltage source is capable of delivering to
the load is given by (5.29) and (5.31) to be
(5.32)
In the case of the practical current source, the maximum deliverable power is
Pi*. = ^ (5 - 33 )
This may be seen from Fig. 5.25 and (5.31) or by (5.32) and Norton's theorem.
We may extend the maximum power transfer theorem to a linear circuit rather
than a single source by means of Thevenin's theorem. That is, the maximum power is
obtained from a linear circuit at a given pair of terminals when the terminals are
loaded by the Thevenin resistance of the circuit. This is obviously true since by
Thevenin's theorem the circuit is equivalent to a practical voltage source with internal
resistance R th .
As an example, we may draw the maximum power from the circuit of Fig. 5.18(a)
if we load terminals a-b with the Thevenin resistance
RL = *,h = 6n
Since /sc — 5 A by (5.22), we may draw the Norton equivalent circuit with the required
RL as shown in Fig. 5.31. The power supplied to the load is given by
_ 90QR L
P
(Rl + 6)
2
Chap. 5 Network Theorems 113
r, = 6a
= 37.5 W
Any other value of RL will result in a lower value of p. For example, if RL 5 ft,
thenwe have/? = 37.19 W, and for RL == 1 Q, we have/? = 37.28 W.
EXERCISES
5.5.1 Find the maximum power that can be transferred from the circuit to the left
of terminals a-b in the figure for Ex. 5.3.1. Ans. 24 W
5.5.2 Show two networks given are equivalent
that the at terminals a-h and find the
power dissipated in the 3-Q resistor in each case.
Ans. (a) 12W, (b)3 W
en. 6Ci
(o)
EXERCISE 5.5.2
5.5.3 Find the maximum power delivered to the load RL in Fig. 5.23 if v
g
and RL >
are fixed and R g is variable. Ans. vljR L when Rg =
114 Network Theorems Chap. 5
PROBLEMS
5.1 Solve Prob. 2.16 using the method applied to Fig. 5.3.
5.3 Solve Prob. 3.13 using the method of Prob. 5.1. (Suggestion: Assume i = mA,
1
etc.)
5.6 Using superposition, find the power delivered to the 3-kfi resistor.
4V
2kA 3k.fl
PROBLEM 5.6
PROBLEM 5.7
4n
PROBLEM 5.8
5.9 Find v in Prob. 4.5 by replacing the network to the left of terminals a-b by its
Thevenin equivalent.
5.10 Find the power delivered to the 4-fi resistor in Prob. 4.9 by replacing the net-
work to the left of terminals a-b by its Norton equivalent.
5.11 Solve Prob. 4.21 by replacing the network to the left of terminals a-b by its
Thevenin equivalent.
5.12 Replace everything except the 6-kfi resistor in Prob. 3.13 by its Thevenin
equivalent and find i.
5.13 Find v by replacing the network to the left of terminals a-b by its Norton
equivalent.
PROBLEM 5.13
5.14 In the figure for Prob. 5.13, replace the network to the right of terminals c-d by
its Thevenin equivalent and find »,.
find v.
116 Network Theorems Chap. 5
PROBLEM 5.15
5.17 Find the maximum power that can be transferred to R by the circuit to the left
3A 2d
PROBLEM 5.17
5.18 In the figure for Prob. 4.22, find the resistance R to replace the 1-Q resistance
between terminals a-b which will draw the maximum power from the rest of the
circuit. Find this maximum power.
5.19 Show that the 1-Q resistance between terminals c-d in the figure for Prob. 4.22
has the correct value to draw the maximum power from the rest of the circuit.
Find this maximum power.
5.20 Solve Prob. 4.23 by assuming v 2 = 1 V and using the proportionality principle.
6
INDEPENDENCE
OF EQUATIONS
An electric network is determined by the type of elements it contains and the manner
in which the elements are connected. We have spent considerable time in the previous
chapters considering the elements themselves and their volt-ampere characteristics. In
this chapter we shall consider the manner in which the network elements are con-
nected, or, as it is called, the network topology. As we shall see, a study of
sometimes
the topology of the network provides us with a systematic way of determining how
many equations are required in the analysis, which ones are independent, and the best
set of equations to select for the most straightforward analysis.
To illustrate the problems involved in the analysis of more complicated networks, let
us consider the circuit of Fig. 6. 1 . The resistors are numbered 2, 9, with values
1 , . . . ,
of resistance, say. R R2
x
, , . . . , R 9 . Suppose we are required to perform a loop analysis,
in which case we need to write a set of independent KVL equations. (We note that the
circuit is nonplanar, and thus we cannot perform a mesh analysis. Anyone doubting
(1, 2, 5, 6, 7, 9), (1, 3, 5, 6, 8, 9), (2, 3, 7, 8), (2, 3, 4, 6, 7, 9), (4, 6, 8, 9), (5, 6, 7, 8),
(1, 3, 4, 6, 7, 8), and (1, 2, 4, 5, 7, 8). That is, resistors 1, 3, 4, and 5 form a loop; 1, 3,
To undertake a loop analysis of Fig. 6.1 we need to know which of these loops are
independent and how many are required. To answer these questions we need consider
117
118 Independence of Equations Chap. 6
only how the elements are connected; it is unimportant what kind of elements are
involved. To facilitate matters, then, we may retain the nodes of the network and
replace its elements by lines. The network topology thus is preserved in a much
simpler configuration.
The configuration of lines and nodes obtained by replacing the elements of a
network by lines is called the graph of the network. The lines of the graph are called
its and the nodes of the graph are, of course, the nodes of the network. As an
branches,
example, the graph of the network of Fig. 6.1 is shown in Fig. 6.2. It has nine branches
and six nodes. (We could consider a node between resistor 9 and v g but inasmuch as ,
this will not affect the number of loops, we have chosen to consider these two series
elements as one element, namely a nonideal voltage source.)
We shall say that a graph is connected if there is a path of one or more branches
between any two nodes. The graph of Fig. 6.2 is evidently connected. An example of a
graph that is not connected is shown in Fig. 6.3. There is, for instance, no path
between nodes a and d. For the present we shall consider only connected graphs.
EXERCISES
6.1.1 Show that the given graph is planar.
EXERCISE 6.1.1
EXERCISE 6.1.2
We define a tree of a graph as a connected portion of the graph that contains all the
nodes but no loops. As an example, Fig. 6.4(b) is a tree of the graph of Fig. 6.4(a). The
tree is connected, has no loops, and contains all the nodes of the graph.
1 20 Independence of Equations Chap. 6
2 4
o o o
Generally, a graph has many trees. The configuration of Fig. 6.4(c) is evidently
another tree of the graph of Fig. 6.4(a), since it satisfies all the requirements. This
particular graph has 24 trees, which the reader may wish to try to discover. It will help
in enumerating the trees to notice that each one has exactly three branches, since it
takes at least three lines to connect four nodes and more than three lines will form a
loop. There are 35 ways to select seven branches three at a time, but 1 1 of these
combinations are not trees.
In the general case, let B be the number of branches and TV be the number of nodes
in a given graph. Then any tree of the graph contains TV nodes and TV — 1 branches.
The number of nodes follows from the definition of a tree, and the number of tree
branches may be established by the following construction argument. Let us build the
tree starting with one branch and the two nodes to which it is connected. Each
additional branch connected to build the tree adds one additional node. The number of
nodes is, therefore, one more than the number of branches, and since there are TV
nodes, there must be TV — 1 branches.
As examples, the graph of Fig. 6.4(a) has TV = 4, and thus the number of tree
branches in both Figs. 6.4(b) and (c) is TV — 1 = 3.
Once the branches of a tree are designated, the remaining branches of the graph are
called links. The number of links graph is then B — (TV — 1) or B — TV + 1. As an
in a
example, the tree of Fig. 6.4(b) is redrawn in Fig. 6.5, with the tree branches shown as
solid lines and the links as dashed lines. The number of links in this case, since B= 7,
is 7 -4+ 1 = 4.
EXERCISES
6.2.1 Find the number of tree branches and links in the graph of Fig. 6.2. Ans. 5, 4
EXERCISE 6.2.2
6.2.3 In the figure for Ex. 6.2.2, let i 2 and / 4 be link currents in elements 2 and 4,
In this section we shall consider the nodal analysis of a circuit whose graph has N
nodes and B branches. A tree of the graph then has N— 1 branches and, of course,
TV — 1 tree branch voltages. There are also B —N+ I link voltages in the circuit.
Let us imagine that all any tree are made zero by short-
the branch voltages of
circuiting the branches (i.e., the branches are replaced by short circuits). Then all the
nodes of the circuit, being in the tree, are at the same potential, and thus the link
voltages are all zero also. Therefore the link voltages depend on the tree branch
voltages, for if a link voltage were independent of the tree voltages, it could not be
forced to zero by short-circuiting the tree branches. On the other hand, if one tree
branch were not short-circuited, there would be one node at a different potential and
thus one tree branch voltage not dependent on the others. We may conclude, there-
fore, that the N — 1 branch voltages of any tree are independent and may be used to
find the link voltages.
Another way to see that any link voltage may be expressed in terms of tree voltages
is adding the link to the tree completes a circuit whose only other elements
to note that
are tree branches. Thus by KVL the link voltage is an algebraic sum of tree branch
voltages.
1 22 Independence of Equations Chap. 6
As an example, the graph of Fig. 6.6 has tree branch voltages v u v 2 and v 3 indi- , ,
cated by the solid lines. The dashed lines are the links whose voltages are v 4 v 5 , and ,
v4 = v3 — v2
v5 = v x
— v2
v6 = Vi - v3
Thus the link voltages may be found from the tree branch voltages.
we have included in the tree the 20-V source. Thus the number of unknowns is reduced
HA
(a)
by one. B> K.VI we may find the link voltages in terms of the tree voltages, with the
results shown in Fig. 6.7(b).
[f we imagine the tree branch (a, b) labeled r, as open, then the tree is separated
into two parts. These two parts are connected by branch (cr, b) and links {a, c), (b, d),
and (</. c), as indicated by the line marked I. Summing the currents across the line in
the direction of the arrow, we have
—
v t + v2 + v t
20
2
v,
l -H=0n
, ,
Repeating the procedure for tree branch (b, c), labeled v 2 , leads to line II and the
equation
— (»i + v i) ~2v 2 + 11 =0
Solving these equations, we have r, = 8 V and v 2 = 1 V. All the link voltages, and
consequently all the link and tree currents, may now be found.
In any circuit analysis procedure where the unknowns are voltages, we need to find
only the N— 1 tree branch voltages which constitute an independent set. This means
that only N — 1 independent voltage equations are required in the analysis. Since any
independent set of equations will suffice, then any independent set of N— 1 voltages
constitutes a solution.
Another independent set of N — 1 voltages, other than the tree branch voltages, is
the set of nondatum node voltages, considered in the nodal method of Chapter 4. To
see this, we note that any nondatum node is in the tree and is connected through tree
branches to the datum node. Thus every nondatum node voltage is an algebraic sum of
tree branch voltages (the tree branches between the nondatum and datum nodes). On
the other hand, every tree branch voltage is the difference between its two node
voltages. In summary, the node voltages may be determined from the tree voltages and
vice versa. Thus the nondatum node voltages are also an independent set.
In the example of Fig. 6.7 we see that if d is the datum node, then the nondatum
node voltages va v b and v c are related to the tree voltages v u v 2 and 20 by
, , ,
va = 20
vb = 20 - »,
vc = 20 — t>i
— v2
Conversely, we have
*>i =%~ vt
v2 = Vb - Vc
20 =v a
The nodal method, in many cases, is easier to apply than the loop method because
the nodes are easy to find. In the loop method, as exemplified by the example of Fig.
6.1, the appropriate loops may be difficult to identify. In the next section we shall
1 24 Independence of Equations Chap. 6
EXERCISES
6.3.1 In the graph of the circuit in the figure for Ex. 4.3. 1
, select the tree of the voltage
sources and the 3-fi resistor. Using the method of this section, write one KCL
equation and determine u lt Ans. —4 V
6.3.2 Select the tree of the voltage source, and the 3- and 6-fi resistors, and use the
method of this section to find v. Ans. 3 V
EXERCISE 6.3.2
6.3.3 Using an appropriate tree and the methods of this section, find v in Prob. 4.18.
(Note: The tree should contain the voltage v and the three sources as well as
one other branch.) Ans. 20 V
As we have seen in the example of Fig. 6.1, it is not always easy to identify the indepen-
dent loops for a loop analysis of a circuit. To develop a systematic means of writing
loop equations, let us consider a general network with B branches and nodes. N
Corresponding to a given tree there are B — N
+ 1 links.
Suppose that all the link currents are made zero by open-circuiting the links
(replacing the links by open circuits). Since the tree contains no loops, then all the tree
branch currents are zero also. The tree currents thus depend on the link currents, i.e.,
they may be expressed in terms of the link currents, for if a tree current were indepen-
dent of the link currents it could not be forced to zero by open-circuiting the links.
Chap. 6 Independence of Equations 1 25
Moreover, if one link is not open-circuited, a loop is left in the graph, and a current
will flow in the link. A link current thus is not dependent on the other link currents. In
summary, the B —N+ l link currents are an independent set, and the loop analysis
of the circuit requires B —N l independent equations.
One systematic way to find B —N+ l independent loops is to start with the tree
and add one of the links. This determines the loop containing that link, since adding
the link to the tree closes a loop. Remove this link and add another link to the tree,
determining a second loop. Continue the process until the B — N -|- l loops are
found. The set is independent because each loop contains a different link.
As an example, let us reconsider the circuit of Fig. 6.1. One tree consists of
branches 1. 5. 7, 8. and 9. with corresponding links 2, 3, 4, and 6, as shown in Fig. 6.8.
Closing the links one at a time results in the four independent loops I, II, III, and IV
shown. Loop I contains link 2 and tree branches 8, 9, and 1 ; loop II contains 3, 7, 9,
and I ; loop III contains 4, 5, 7, and 9; and loop IV contains 6, 5, 7, and 8. These four
loops are sufficient for performing a loop analysis.
To illustrate the use of link currents in circuit analysis, let us return to the example
of Fig. 6.7(a). The graph is redrawn in Fig. 6.9, showing the link currents and /,, i 2 ,
1 1 A. We have chosen the current source as a link because the link currents are an
independent set. This reduces the number of unknowns by one. Generally, for this
reason, one should place voltage sources in the tree and current sources in the links.
The tree branch currents, as in the general case, may be found from the link
currents, as shown in Fig. 6.9. Closing the links labeled /', and i2 forms loops 1 and 2, as
indicated. Applying KVL to these loops yields, from Figs. 6.7(a) and 6.9,
2/, - 20 + /J
— i2 + 11 =
h -F l: 11=0
i, - i,+ II loop 2
loop I
FIGURE 6.8 Circuits of Fig. 6.1 FIGURE 6.9 Graph of Fig. 6.7(a)
1 26 Independence of Equations Chap. 6
Since the link current 1 1 A is known, we needed only two loops involving link
currents i, and i2 . Incidentally, in this simple example the links were chosen so that the
link currents are also mesh currents. This is, of course, not the case in general.
The results obtained thus far in this chapter are valid for general networks, which
may be either planar or nonplanar. In the special case of planar networks, as we saw in
Chapter 4, a mesh analysis is possible. In the circuits of that chapter the mesh cur-
rents were independent and were sufficient in number to perform the analysis. We
shall now show that this is the case in general for planar networks.
Let us begin by taking apart the planar circuit with meshes and reconstructing it M
one mesh at a time. The first mesh in the reconstruction has the same number, say k u
of nodes and branches, for the first branch has two nodes, each additional branch
adds one new node, and the last branch adds no nodes since it is tied back to a node of
the first branch. This is illustrated by the graph of four meshes in Fig. 6.10(a). The
first mesh constructed, shown in Fig. 6.10(b), has the same number of branches and
D o
T 9 9
o 6 o
9 9
o o 9
6 -o o
(d) (e)
After the first mesh, each subsequent mesh is formed by connecting branches and
nodes to previous meshes. Each time the number of nodes added is one less than the
number of branches, because each added branch adds one node, except for the last
branch, which is connected to a node of a previously added mesh. This process is
illustrated in Figs. 6.10(c), (d), and (e).
Chap. 6 Independence of Equations 1 27
Similarly, the third mesh adds k branches and k — nodes, and so forth. The last
3 2
l
mesh, the .Uth. adds k M branches and k M — nodes. If in the completed graph the l
k x
+ k2 + ... + kM -(M- \) =N
which by (6.1) becomes
B - {M - 1) =N
Solving for the number of meshes, we have
M=B-N+ I (6.3)
pletely describe a planar network. They are the same number as the independent set
in
of link currents, and they are independent since each new mesh contains at least one
branch not in the previous meshes.
EXERCISE
6.4.1 Show that (6.3) holds for the circuits of Probs. 4.19, 4.20, and 4.22.
Thus if the tree branch voltage method is used, there will be only two unknowns, v x
and v 2 requiring two equations. There are five link currents in the graph, one of which
.
is known (the 6-A source) and another, the 2v source, which may be expressed in l
1 28 Independence of Equations Chap. 6
IOV
IOV
(a)
(b)
terms of the tree branch current i be , and subsequently in terms of other link currents.
Thus if the link current method is used, we must have three equations. Accordingly, we
shall analyze the circuit using tree branch voltages.
To write thetwo necessary equations we first imagine tree branch (b, e) opened,
which separates the tree into two parts. These parts are connected by branch (b, e) and
links (b, c), (c, d), and (a, e) of the 6-A source and link {a, e) of the ^-Q resistor. These
elements are identified by line I. A similar procedure using branch (a, d) leads to line
II. Applying KCL to the currents crossing lines I and II in the direction indicated
yields
and
2v.dc 2v 1
3v 2 =0 (6.5)
v bc = Vi — 3v 2
v dc = v, - 2v 2 10
Vac = »! — 10
Substituting these values into (6.4) and (6.5) and solving for the tree branch voltages,
we have
i>, = -11 V, v2 = -20 V
We may note that nodal analysis is equally easy to apply. The node voltages may
Chap. 6 Independence of Equations 1 29
be expressed in terms of the two unknowns r, and ?\, and thus only two nodal
equations are required. We shall leave the details to the problems (Prob. 6.4).
EXERCISES
6.5.1 Using the graph described for the figure for Ex. 6.3.2, write one KVL equation
and find using the method of Sec. 6.4.
/', Ans. 3 A
6.5.2 Using an appropriate tree for the graph of Prob. 4. 1 6, find the current i flowing
to the right in the 4-fi resistor. (An appropriate tree should not contain the
current sources or the current i. Thus, using the methods of Sec. 6.4, only one
KVL equation is required.) Ans. —9.5 A
6.5.3 Solve Prob. 4.9 using the method of Sec. 6.4. Ans. 4 W
PROBLEMS
6.1 Find a tree, if possible, that contains all the voltage sources and the branches
whose voltages control dependent sources but does not contain current sources
or branches whose currents control dependent sources. Use this tree with an
appropriate graph theory method to find v x
.
PROBLEM 6.1
6.2 Select a tree as described in Prob. 6.1 and use an appropriate graph theory
method to find y t .
1 30 Independence of Equations Chap. 6
2i V
I5V
PROBLEM 6.2
6.3 Solve Prob. 4.21 selecting an appropriate tree and using graph theory methods.
6.5 In the figure shown, all the resistances are 1 Q, element w is also a 1-Q resistor,
elements x and y are independent 1-A current sources directed upward, and
element z is an independent 3-A current source directed to the left. Selecting an
appropriate tree and using graph theory methods, find i. (Note that this circuit
is similar to that of Fig. 6.1 and is thus nonplanar. However, only one loop
equation is required in this case.)
PROBLEM 6.5
6.6 Find v in Prob. 6.5 using graph theory methods if element x is a 4-V source
with positive terminal at the top, y is a 2-V source with positive terminal at the
bottom, z is a 6V source with positive terminal at the left, and w is a 4-V source
with positive terminal at the left.
6.7 The figure for Ex. 6.2.2 and Figs, (a) and (b) shown are examples of graphs of
ladder networks. There is a theorem that states that the number of trees in a
Chap. 6 Independence of Equations 1 31
a2 = a + a t
= 2, o3 = a x
+ a2 = 3, a4 = a 2 + a 3 = 5, etc. That is, except
for a and a u each Fibonacci number is obtained by adding the previous two.
Verify that the theorem holds for the ladder graphs shown and also for the graph
in the figure for Ex. 6.2.2.
(a) (b)
PROBLEM 6.7
6.8 The graphs shown two basic nonplanar graphs [the one in (b)
in (a) and (b) are
is that in the figure for Prob. 6.5 redrawn]. Branch a-b in (a) is an ideal 6-V
voltage source with its positive terminal at the top and in (b) is a 4-A ideal cur-
rent source directed upward. All the other branches in both figures are 1-Q
resistors. Find i shown in each figure using graph theory methods.
(a) (b)
PROBLEM 6.8
6.9 Find y a and v 2 in Fig. 6.11 (a) using link currents as the unknowns.
Up to now we have considered only resistive circuits, that is, circuits containing
resistors and sources. The terminal characteristics of these elements are simple
algebraic equations which lead to circuit equations that are algebraic. In this chapter
we shall introduce two important dynamic circuit elements, the capacitor and the
inductor, whose terminal equations are differential equations rather than algebraic
equations. These elements are referred to as dynamic because, in the ideal case, they
store energy which can be retrieved at some later time. Another term which is used, for
this reason, is storage elements.
We shall first describe the property of capacitance and discuss the mathematical
model of an ideal device. The terminal characteristics and energy relations will then be
given, followed by derivations for parallel and series connections of two or more
capacitors. We shall then repeat this procedure for the inductor.
The chapter will be concluded with a discussion of practical capacitors and
inductors and their equivalent circuits.
7.1 CAPACITORS
A capacitor is a two-terminal device that consists of two conducting bodies that are
separated by a nonconducting material. Such a nonconducting material is known as a
As a result of the dielectric, charges are not permitted to move from one
dielectric.
conducting body to the other within the device. They must therefore be transported
between the conducting bodies via external circuitry connected to the terminals of the
capacitor. One very simple type, called a parallel-plate capacitor, is shown in Fig. 7.1.
The conducting bodies are flat, rectangular conductors that are separated by the
dielectric material.
132
Chap. 7 Energy-Storage, Elements 1 33
To describe the charge-voltage relationship for the device, let us transfer charge
from one plate to the other. Suppose, for instance, by means of some external circuit,
that we take a small number of electrons, say — Aq, from the upper plate to the lower
plate. This, of course, leaves a charge of Ar/ on the top plate and deposits a charge of
I
— Aq on the bottom plate. Since moving these charges requires the separation of unlike
charges (recall that unlike charges attract one another), a small amount of work is
performed, and the top plate is raised to a potential of say -\-Av with respect to the
bottom plate.
Each increment of charge — Aq that we transfer increases the potential difference
between the plates by Ar. Therefore the potential difference between the plates is
proportional to the charge being transferred. This suggests that a change in the
terminal voltage by an amount Av causes a corresponding change in the charge on the
upper plate by an amount Aq. Thus the charge is proportional to the potential differ-
ence, and we may write sr
q = Cv fatA (7.1)
Vo i
o-
dielectric +
^c
terminal. The movement of this charge causes the upper terminal to become more
positive than the lower one by an amount Av. Hence the current-voltage convention of
Fig. 7.2 is satisfied. Obviously, if v is assigned so that the lower terminal is positive or
if the current is assumed to enter the lower terminal, but not both, then a minus sign
must be used in the right-hand side of (7.2). We recall that this is also necessary in the
case of a resistor.
In (7.2) we see that if v is constant, then the current i is zero. Therefore a capacitor
acts like an open circuit to a dc voltage. On the other hand, the more rapidly v changes,
the larger is the current flowing through its terminals. Consider, for example, a voltage
which increases linearly from to 1 V in a"
1
s, given by
v = 0, t<
= at, 0<t<a~ l
= 1, t>a-*
If this voltage is applied to the terminals of a 1-F capacitor (an unusually large value
which is convenient for illustrative purposes), the resulting current is
i = 0, / <
= a, < < t a- 1
= 0, txr 1
Plots of v and i areshown in Fig. 7.3. We see that i is zero when v is constant and
that it is equal to a when v increases linearly. If a is made larger, then v changes more
rapidly and i increases. It is apparent that if a
-1
= (a is infinite), v changes abruptly
(in zero time) from to 1 V.
In general any abrupt or instantaneous changes in voltage, such as in the above
example, require that an infinite current flow through the capacitor. An infinite
current, however, requires that an infinite power exist at the capacitor terminals, which
/\
//
1
l/a t
= ±f'f<ft
C
+ t</ ) yJ (7.3)
where v(t ) = q(t )/C is the voltage on C at time f . In this equation, the integral term
represents the voltage that accumulates on the capacitor in the interval from t to /,
whereas r(/ ) is that which accumulates from -co to t The voltage v{— oo), of .
v(t) = -L \
idt
In applying this result, we obviously are obtaining the area associated with a plot
of/ from — oc to f. In Fig. 7.3, for example, since v(— oo) = 0, we have
v = -^ |
(0) dt + v(- >o) = 0, t <
at the same time, unlike the case for the resistor. In fact, inspection of Fig. 7.3 reveals
that the current can be discontinuous even though the voltage is continuous, as
stated previously.
1 36 Energy-Storage Elements Chap. 7
EXERCISES
7.1.1 If a 100-//F capacitor has v(t) = f(t) V, as shown, find (a) /(-3), (b) i(-l), (c)
i(1.5), and (d) /(2.5). (e) Sketch i(f) for all t.
EXERCISE 7.1.1
7.1.2 If a 1000- //F capacitor has i(t) = f(t) mA, where f(t) is given in Ex. 7.1.1, find
(a) u(-2), (b) v(0), and (c) v(2). (d) Sketch v(t) for all t. [Note: v(— 4) = 0.]
Ans. (a) 1 V, (b) 2 V, (c) 0.5 V
7.1.3 A 1-juF capacitor has a terminal voltage of 100 cos 1000/ V. Find /(?)
Ans. -0.1 sin 1000/ A
7.1.4 A 10-//F capacitor has a terminal current of lOe -100 ' mA. If v(0) = — 10 V,
find v(t) for / > 0. Ans. -lOe" 100 '
V
between the capacitor plates. These charges have electrical forces acting on them. An
electric field, a basic quantity in electromagnetic theory, is defined as the force acting
on a unit positive charge. Thus the forces acting on the charges within the capacitor
can be considered to result from an electric field. It is for this reason that the energy
stored or accumulated in a capacitor is said to be stored in the electric field.
The energy stored in a capacitor, from (1.6) and (7.2), is given by
=C [ vdv = \Cv\t)
Chap. 7 Energy-Storage Elements 137
«e>-i*P> (7.5)
The ideal capacitor, unlike the resistor, cannot dissipate any energy. The energy
which is stored in the device can thus be recovered. Consider, for instance, a l-F
capacitor which has a voltage of 10 V. The energy stored is
if, = \Cv 2 = 50 J
Suppose the capacitor is not connected in a circuit; then no current can flow, and
the charge, voltage, and energy remain constant. If we now connect a resistor across the
capacitor, a current flows until all the energy (50 J) is absorbed as heat by the resistor
and the voltage across the combination is zero. The analysis of such a network is
found in the next chapter.
As has been pointed out earlier, the voltage on a capacitor is a continuous func-
tion. Thus by (7.4) we see that the energy stored in a capacitor is also continuous. This
is not surprising since otherwise energy would have to be transported from one place
t =
AW-
+
C ZZ
pair of terminals from an open circuit to a short circuit, or vice versa, in zero time.) To
discuss the effect of the switching action we first need to consider two different types of
time / = 0. We shall denote / = 0" as the time just before the switching action and
/ —
+
as the time just after the switching action. Theoretically, of course, no time has
elapsed between 0" and + but the two times represent radically different states of the
,
circuit. Thus v c (0~) is the voltage on the capacitor just before the switch is moved and
approaches zero through positive (/ > 0) values. The same notation applies to v l
across the resistor J?,.
Suppose that in Fig. 7.4 we have V = 6 V and v c (0~) = 4 V. Just prior to the
switching action (t = 0") we have v^O') = V — v c (0~) = 2 V. Immediately after the
switch is opened we have V!(0 + ) = 0, since no current is flowing in i?,. However, since
vc is continuous we have
v c (0+) = vc (0~) = 4V
Thus the voltage on the resistor has changed abruptly, but that on the capacitor has
not.
Obviously, one could consider circuits on paper in which capacitor voltages are
forced to change abruptly. For example, if two capacitors having different voltages
are suddenly connected in parallel by a switching action, their resulting common
voltage cannot be the same as both their previous, different voltages. We shall consider
such singular circuits in Sec. 7.9, where we shall see that stored energy has appeared to
change abruptly. The apparent change cannot be accounted for in the lumped circuit
models we are using, but it is a remarkable fact that the lumped models are valid
before and after (though not during) the switching action. Physical circuits, however,
have resistance associated with the capacitor (such as in the leads and the dielectric)
which precludes the infinite currents that must accompany discontinuous capacitor
voltages. These are the type circuits we shall be concerned with, in general.
EXERCISES
7.2.1 A 1-//F capacitor is charged to 100 V. Find the charge and energy.
Arts. 10" 4 C, 5 mJ
7.2.2 In Fig. 7.4, let C= £ F, J?, = R2 = 3 ft, and V= 9 V. If the current in R 2 at
/ = 0" is 1 A directed downward, find at t = 0" and at / = + (a) the charge
on the capacitor, (b) the current in R x
directed to the right, and (c) the current
inC directed downward. Am. (a) 1, 1 C, (b) 2, A, (c) 1, —1 A
In this section we shall determine the equivalent capacitance for series and parallel
connections of capacitors. As we shall see, equivalent capacitance is in direct analogy
with equivalent conductance.
1/
Chap. 7 Energy-Storage Elements 1 39
Let us first consider the series connection of N capacitors, as shown in Fig. 7.5(a).
Applying K.VL. we find
v = «, + v2 + . . . + vN (7.6)
i
+ ' _ + < _
=^H{ If-
Q)
(a) (b)
= 1 f 1 f
v -^ \ idt + w,(/ ) + £- / <fc + w 2 (/ )
+ ... + -J- \
idt + V N (t )
1 1
=
(^ + gr + • • • + ^)
J'/
<* + »i(f„) + ^('o) + • • • + tvOo)
or, by (7.6),
v = ^ I / dt + K'o)
Suppose we require that the circuit of Fig. 7.5(b) be an equivalent circuit for that
of Fig. 7.5(a). Comparing the last two equations, we see that
As an example of the utility of (7.6) and (7.7), consider the series connection of
1- and ^-F capacitors having initial voltages of 4 and 6 V, respectively. Then
&- 1
or
C. = 0.25 F
and
v(t ) = 4 + 6= 10 V
/ = /, + iz + . . . + i N
^ dv . „ dv ,
CN
dv
dt
dv dv
(C, + c2 + . . . + CN )
di (I
c-)
di
(a) (b)
If we now require that this circuit be an equivalent circuit for that of Fig. 7.6(a), the
above equations give
C, = C, +C + 2 . . . + C„ = £ C„ (7.8)
n= 1
Thus the equivalent capacitance of TV parallel capacitors is simply the sum of the
individual capacitances. An initial voltage, of course, would be equal to that which is
EXERCISES
7.3.1 Find the equivalent capacitance and the initial voltage. Ans. 0.1 F, —2 V
2V 3V
'/
2 F '/3F
'/5 F 7ZZ 7 V
EXERCISE 7.3.1
o 1 p 1 1
-
v(t) CIF ~2F 3F
o 1
i -4 i
EXERCISE 7.3.2
6(iF
EXERCISE 7.3.3
1 42 Energy-Storage Elements Chap. 7
7.4 INDUCTORS
In the previous sections, we found that the electrical characteristics of the capacitor
are the result of forces that exist between electric charges. Just as static charges exert
forcesupon one another, it is found that moving charges, or currents, also influence
one another. The force which is experienced by two neighboring current-carrying
wires was experimentally determined by Ampere in the early nineteenth century.
These forces can be characterized by the existence of a magnetic field. The magnetic
field, in turn, can be thought of in terms of a magnetic flux that forms closed loops
about electric currents. The origin of the flux, of course, is the electric currents. The
study of magnetic fields, like that of electric fields, comes in a later course on electro-
magnetic theory.
An inductor is a two-terminal device that consists of a coiled conducting wire. A
current flowing through the device produces a magnetic flux which forms closed <f>
loops encircling the coils making up the inductor, as shown by the simple model of
Fig. 7.7. Suppose that the coil contains N
turns and that the flux passes through each <f>
turn. In this case, the total flux linked by the turns of the coil is N
= N<f>
This total flux is commonly referred to as the flux linkage. The unit of magnetic flux is
the weber ( Wb), named for the German physicist Wilhelm Weber (1804-1891).
In a linear inductor, the flux linkage is directly proportional to the current flowing
through the device. Therefore, we may write
/ \ X=Li (7.9)
where L, the constant of proportionality, is the inductance in webers per ampere. The
unit of 1 Wb/A is known as the henry (H), named for the American physicist Joseph
Henry (1797-1878).
In (7.9) we see that an increase in i produces a corresponding increase in A. This
increase in A produces a voltage in the JV-turn coil. The fact that voltages occur with
changing magnetic flux was first discovered by Henry. Henry, however, repeating the
mistake of Cavendish with the resistor, failed to publish his findings. As a result, the
famous Michael Faraday is credited with discovering the law of electromagnetic
induction. This law states that the voltage is equal to the time rate of change of the
total magnetic flux. In mathematical form, the law is
dt
di
V V = ,
L-r- (7.10)
dt
Chap. 7 Energy-Storage Elements 1 43
defined by
i = 1, t<0
= 1
-
bt, 0<t<b~ l
= 0, t> b~ l
A 1-H inductor having this terminal current has a terminal voltage given by
v = 0, t <
= -b, 0<t<b- 1
= 0, t>b~ l
Plots of/ and v for this case are shown in Fig. 7.9. We see that v is zero when / is
\ "b '
'/b
FIGURE 7.7 Simple model of an FIGURE 7.8 Circuit FIGURE 7.9 Current and voltage wave-
inductor symbol for the inductor forms for a 1-H inductor
1 44 Energy-Storage Elements Chap. 7
case of the capacitor, this requires that an infinite power exist at the terminals of the
inductor, a physical impossibility. Thus instantaneous changes in the current through
an inductor are not possible. We observe that the current is continuous even though
the voltage may be discontinuous.
An alternative statement concerning abrupt changes of the currents flowing in
circuits containing more than one inductor is that the total flux linkage cannot change
instantaneously. That is,Lu L2 L N the sum
for a circuit containing inductors , . . . , ,
A[ + X2 + . . . +
X N cannot change instantaneously. If we compare (7.1) and (7.9),
we see that the flux linkage in an inductor is analogous to the charge on a capacitor.
Thus the sum of the flux linkages given above (conservation of flux linkage) is ana-
logous to conservation of charge. An example employing the conservation of flux
linkage is given in Sec. 7.9.
Let us now find the current i(t) in terms of the voltage v(t). Integrating (7.10) from
time t to / and solving for i(t), we have
In this equation, the integral term represents the current buildup from time t to t,
whereas i(t ) is the current at t . Obviously, /'(/ ) is the current which accumulates
from t = — oo to t , where /(— °°) = 0. Thus an alternative expression is
i(')
H>
= xj V{t)dt
In the application of (7.11), we are obtaining the net area under the graph of v
from — co to /, since /(/„) represents the area from -co to t . In Fig. 7.9, for instance,
since /(0) = 1, we have, for L= 1 H,
In the above example, we see that v and /Just as in the case of the capacitor, do not
necessarily have the same variation in time. Inspection of Fig. 7.9, for example, shows
Chap. 7 Energy-Storage Elements 145
that the voltage can be discontinuous even though the current is continuous, as
previously mentioned.
EXERCISES
7.4.1 If a 100-mH inductor has /(/) =/(/)mA, where /(0 is given in Ex. 7.1.1, find
(a) K-3),(b) K-l), (c) r(1.5), and (d) i>(2.5). (e) Sketch v{t) for all t.
7.4.2 If a 10-mH inductor has v(t) =f(t) mV, where f(t) is given in Ex. 7.1.1, find (a)
i(-2), (b) ;'(0), and (c) i(2). (d) Sketch i(t) for all t. [Note: /(— 4) = 0.]
Ans. (a) 0.1 A, (b) 0.2 A, (c) 50 mA
7.4.3 A 1-mH inductor has a terminal current of 10 sin 100/ A. Find v(t).
'
= L\' i di = %Li\t)
J-oo f=-oo
Inspection of this equation reveals that w L (t) > 0. Therefore, from (1.7), we see
that the inductor is a passive circuit element.
The ideal inductor, like the ideal capacitor, does not dissipate any power. There-
fore the energy stored in the inductor can be recovered. Consider, for example, a 2-H
r (
1 46 Energy-Storage Elements
Ui- / 2- U* I
C/73P. 7
^. V wL = ill = 2
25 J
the current is zero. Solutions of circuits of this type are found in the next chapter.
Since inductor currents are continuous, it follows that the energy stored in an
inductor, like that stored in a capacitor, is also continuous. To illustrate this let us
consider the circuit of Fig. 7.10, which contains a switch that is closed at t = 0, as
indicated. Suppose that /
L (0~)= 2A and / = 3 A. Then by KCL, i x
{0)~) = 3 —2=
1 A. After the switch is closed (t = +
), we have /](0
+
) = since a short circuit is
/t (0+) = /L (0-) = 2A
Thus the resistor current has changed abruptly but the inductor current has not.
current
apparent discontinuity in the energy stored in the inductors cannot be accounted for in
the lumped circuit model. However, lumped circuit theory is valid before and after
(though not during) the switching action. Physical circuits having inductors contain
associated resistance which does not permit the infinite inductor voltages that must
accompany abrupt changes in inductor currents. We shall be concerned primarily
with circuits of this type.
EXERCISES
7.5.1 Derive an expression for the energy stored in an inductor in terms of the flux
2
linkage X and the inductance L. Ans. 2. /2L
7.5.2 A 10-mH inductor has a current of 100 mA. Find the flux linkage and the energy.
Ans. 1 mWb, 50 juJ
Chap. 7 Energy-Storage Elements 147
In this section we shall determine the equivalent inductance for series and parallel
connections of inductors. Let us first consider a series connection of TV inductors, as
shown in Fig. 7.11(a). Applying KVL, we see that
v =v +v +
t 2 . . . + vN
di di di
v=L t
+ L ** +
.
T ,
--- +
,
Lr
>Tt »li
= (L,+L + ...+L N
)f 2
(
dl
•fa) dt
(a) (b)
Ls
dt
If we now require that this circuit be an equivalent circuit for the series connection,
then the above equations yield -.
V
L s
— Lj -r Li — . . . + LN — E 2-i Ln (7.13)
n= 1
Therefore the equivalent inductance of N series inductors is simply the sum of the
individual inductances. In addition, an initial current would clearly be equal to that
flowing in the series connection.
Let us now consider the parallel connection of N inductors, as shown in Fig.
7.12(a). Application of KCL gives
* =h+ it + . . . + i N (7.14)
148 Energy-Storage Elements Chap. 7
(a) (b)
i = i- f V dt + /,(/„) + £" f
« * + ^o) + + fifr + /*(/„)
= + zb)
(r + r+ 2
• •
•
/
v
*+ /l( ' o) + '
2( ' o) + • • • + /;v( ' o)
or, by (7.14),
i =— v *+ /(* )
where ;(/ ) i s the current in Lp at? = /„. If this circuit is an equivalent network for the
parallel connection, then the above equations require that the equivalent parallel
s—Liii
inductance be given by
L L = + + ...
i^i->,
+ 7L = i:~
dt -
(7.15)
7 7 7
As an example of the application of (7.14) and (7.15), suppose we have two parallel
inductors of 2 and 3 H carrying initial currents of 2 and 1 A, respectively. The parallel
combination could be replaced by a single inductance Lp satisfying
1 1 1
*<, ' 2
or
^ = |H
carrying an initial current of
i{t ) = 2 + 1 = 3A
Chap. 7 Energy-Storage Elements 1 49
EXERCISES
2H 4H
-/nsin-
i(0) =IA
v(t) 3H
EXERCISE 7.6.1
7.6.2 Find the equivalent inductance and the initial current. The currents shown are
initial values. Ans. 0.1 H, —2 A
''5H
EXERCISE 7.6.2
EXERCISE 7.6.3
1 50 Energy-Storage Elements Chap. 7
used, the oxide will be reduced, and heavy conduction will occur between the plates.
Practical inductors, like practical capacitors, usually dissipate a small amount of
power. This dissipation results from ohmic losses associated with the wire making up
the inductor coil and core losses due to induced currents arising in the core on which
the coil wound. An equivalent circuit for an inductor can be realized by placing a
is
resistance in series with an ideal inductor, as shown in Fig. 7.14, where R L represents
the ohmic losses and L the inductance.
Inductors are available with values ranging from 1 //H to 100 H. Large inductance
values are obtained by employing many turns and ferrous (iron) core materials;
hence as the inductance increases, the series resistance generally increases.
Chap. 7 Energy-Storage Elements 1 51
Like the resistor and the operational amplifier, the capacitor can be fabricated in
integrated-circuit form. However, attempts at integrating the inductor have not been
very successful because of geometry constraints and because semiconductors do not
exhibit the necessary magnetic properties. For this reason, in many applications,
circuits are designed using only resistors, capacitors, and electronic devices, such as
op amps.
EXERCISE
7.7.1 Mylar capacitors have a resistance-capacitance product of 10 5 Q-F. Find the
equivalent parallel resistor in Fig. 7.13 for the following capacitors: (a) 100 pF,
(b) 0.1 n¥, and (c) 1 //F. Ans. (a) 10 15 Q, (b) 10 12 fi, (c) 10 11 Q
Let us now determine the dual relationships for the capacitor and the inductor. This
is easily done by considering the current-voltage relations of (7.2) and (7.10) for the
elements. Repeating these equations for convenience, we have
i = C dv (7.16)
dt
and
r d
dt
(7.17)
Voltage Current
Charge Flux
Resistance Conductance
Inductance Capacitance
Short circuit Open circuit
Impedance Admittance*
Nonreference node Mesh
Reference node Outer mesh
Tree branch Link*
Series Parallel
KCL KVL
* Tree branch and link were considered in Chapter 6,
We are now able to construct dual circuits for networks containing the dual
quantities listed in Table 7.1. Consider, for example, the two-mesh network of Fig.
7.15(a). Using the geometric method of Sec. 4.7, the dual circuit is shown dashed in
Fig. 7.15(a) and is redrawn in Fig. 7.15(b). The solutions for the currents and voltages
(a)
(b)
EXERCISE
7.8.1 Construct dual circuits for the networks of (a) Fig. 7.4, (b) Fig. 7.10, (c) Fig.
A circuit inwhich a switching action takes place that appears to produce discon-
tinuities in capacitor voltages or inductor currents is sometimes called a singular
circuit. In this section we shall consider two such circuits, one containing capacitors
tively, prior to the closing of the switch. That is, ^(0") = V and v 2 (0~) = V. We 1
shall now determine n'i(0 ) and vv,(0 + ), the energies stored in C, and C 2 at / =
+ +
.
w 1
(p-) = $C v\(0-) = $J1
and
w 2 (0-) = $C2 vl(0-) = 03
qi (o-) = c^co-) = i c
and
q 2 (0~) =C 2v2
(0-) = C
After the switch closes, we see from KVL that
i<,(0
+
) = w2 (0 + )
1 54 Energy-Storage Elements Chap. 7
?.(0
+
) = ? 2 (0
+
)
In addition, when the switch is closed, conservation of charge requires that the total
charge remain constant; hence
?,(0
+
) + q z (0
+
) = 2^,(0
+
) d 2q 2 (0+) = C l
or
?>(0
+
) = ? 2 (0
+
) = iC
Substituting into (7.1), we find
v (o
l
+
) = v 2 (o + ) = ^\
Therefore the energy stored in each capacitor at / = +
is
Let us now compare the energy in the system before and after closing the switch.
At / = 0",
Wi(0-) + w2 (0-) = $J
At t = +
,
w:(0 + ) + w 2 (0+) = \ J
We know that capacitors do not dissipate power. What, then, has happened to the
\ from = 0" to =
+
J t ? Looking back over our work, we see that v
t changes x
the voltage are not possible. Therefore during the infinitesimal time from = 0" to t
t = +
, our mathematical model is not valid. In what has happened is the
reality,
following. When the switch closes at time t = 0, a large current is produced as charges
are transferredfrom Cj to C 2 This rapidly changing current gives rise to an electro-
.
magnetic wave which radiates \ J of energy. The voltage v changes in a short, but l
nonzero, time from 1 to \ V. Our network during this interval does not behave as a
lumped-parameter circuit, and concepts from electromagnetic theory (a later course)
are required for the solution we have described.
Although our circuit model is not valid at the instant the switch closes, the solutions
for the voltages and energies before and after the closing of the switch are correct. This
is due entirely to the fact that the total charge did not change during this time interval.
As pointed out earlier, most circuit models do not permit an infinite current in a
capacitance. Physical circuits normally have a finite value of resistance and inductance
which limit such currents. As a result, the capacitor voltages and energies are con-
Chap. 7 Energy-Storage Elements 1 55
tinuous functions. If. for example, a series resistance is included in the circuit of Fig.
7. 16. the voltage on each capacitor is continuous. That is,
u,(0-) = v,(0
+
)
and
v 2 (0-) = v 2 (0 + )
FIGURE 7.16 Circuit containing two FIGURE 7.17 Circuit containing two inductors
capacitors which are switched at time t = which are switched at time t =
The energy stored in the network prior to the closing of the switch is
h^(O-) = iLj/ftO-) = 1 J
and
H' 2 (0-) = jL 2/ 2
(0-) = 0J
A lv0-) = £,i,(0-) = 2 Wb
and
A,(0-) =L 2/ 2
(0-) = Wb
After the switch is closed, we see from KCL that
i.CO*) = /
2 (o+)
Also, when the switch is closed, conservation of flux linkage requires that the total
flux remain constant; hence
= (L, + L )/,(0 + 2 )
1 56 Energy-Storage Elements Chap. 7
or
2(1) + 1(0) = 3/ t (0 )
+
Therefore
/,(0
+
) = / 2 (0
+
) = fA
Thus the energy stored in each inductor at / = +
is
w I (0
+
) = iL 1 iK0
+
) = |J y
and
w 2 (o+) = ^L 2 m +
) = i j
If we now compare the total energy stored in the network, we see at / = 0" that
h',(0-) + W2 (oA = i j
and at / = +
that
h' 1 (0
+
) +w 2 (0
+
) = fJ
which indicates that % J has been lost by the circuit even though ideal inductors can
dissipate no power.
Looking back over the problem, we see that i (t) changes abruptly from 1 to \ A t
at t— 0. We know, however, that abrupt changes in the current are not possible.
Therefore during the infinitesimal time from t = 0" to t — + our mathematical model ,
is once again not valid. As pointed out previously, a rapidly changing current gives
rise to an electromagnetic wave which radiates energy. In this case ^ J is radiated, and
the current changes from 1 to ^ A in a short, but nonzero, time. Our circuit, of
course, does not behave as a lumped-parameter network during this interval of time.
As in the case of a capacitive circuit having infinite currents, most inductive circuit
models do not permit infinite voltages to occur across an inductor as a result of abrupt
currents. As discussed in the case of the capacitor, physical circuits having inductors
contain a finite value of resistance and capacitance which limit such voltages. The
currents and energies in these circuits are continuous functions. If, for instance, a
parallel resistance is included in Fig. 7.17, the currents are continuous at / = 0.
EXERCISES
7.9.1 If in Fig. 7.16 C 2 is changed to i F and ^(O - ) to 1 V, find the voltage and the
total stored energy at t = +
. Ans. |V, jJ
7.9.2 If L z = 3 H in Fig. 7.17, find the current and the total energy after the switch
_
is closed if i,(0 ) = 1 A. Ans. \ A, \ J
Chap. 7 Energy-Storage Elements 1 57
PROBLEMS
7.1 Find the charge residing on each plate of a 2-//F capacitor that is charged to
100 V. If the same charge resides on a 1-//F capacitor, what is the voltage?
7.2 Determine the voltage required to store 50 fiC on a 0.25-//F capacitor. What
mA to deliver this charge?
time will be required for a constant current of 25
7.3 The voltage on a 0.01 -//F capacitor increases linearly from V at / = to 100 V
at / = 10 ms. It then decreases linearly to V at / — 30 ms. Find (a) </(5), (b)
/'(15), and (c)/>(15) (arguments in milliseconds), (d) Sketch q{t), i{t), and p(t).
7.4 Determine the current flowing in a 0.1 -jnF capacitor having the following ter-
minal voltages: (a) 10, (b) 100/, (c) 50^"'°', (d) 150 cos 100/, and (e) 100c"' sin /
V.
7.6 Find the work required to charge a 0.1 -/zF capacitor to 100 V.
t =
I F 2F v (t)
2
PROBLEM 7.8
7.9 What are the maximum and minimum values of capacitance that can be obtained
from 10 1-yWF capacitors?
7.10 The capacitances shown are all in microfarads. Find Ceq at terminals a-b (see
figure below).
7.11 Derive an equation for current division between two parallel capacitors.
7.12 Derive an equation for voltage division between two series capacitors.
7.13 Determine the flux linkage of a 100-mH inductor carrying a current of 0.1 A.
1
1 i
\(
it K
24 12
I2P ^20 ~6
4 8
•^ 4
\(
\\
,
'
If
\\
1
2 ^
2^: ~4
1 1
i i 1 i
PROBLEM 7.10
7.15 Determine the terminal voltage for a 10-mH inductor having the following
terminal currents: (a) 10 A, (b) 50t 2 A, (c) 10 sin 377? A, and (d) 100/e'' A.
7.17 Determine the terminal current for a 10-mH inductor having a terminal volt-
age of lOe" 100 '
V if i(0) =0A.
7.18 Find the work required to establish a current of 100 mA in a 10-mH inductor.
(0") = i A 3H 3 |
i
2
(0") = 2A
PROBLEM 7.20
7.21 Find the maximum and minimum values of inductance which can be obtained
from five 10-mH inductors.
7.23 Derive an equation for voltage division between two series inductors.
7.24 Derive an equation for current division between two parallel inductors.
*7.25 A 400- and a 600-pF ceramic capacitor are connected in parallel. The resistance-
capacitance product for a ceramic capacitor is 10 3 fi-F. What is the equivalent
capacitance and parallel resistance for the combination?
Chap. 7 Energy-Storage Elements 159
PROBLEM 7.22
7.26 Determine a dual circuit for the network of (a) Fig. 7.13, (b) the figure for
"7.27 (a) Make R = in Prob. 7.8 and find the energy lost during the switching
action due to radiation, (b) Replace R by an open circuit in Prob. 7.20 and
determine ^(O*).
8
SIMPLE RC AND Rl CIRCUITS
In this chapter we shall consider simple circuits containing resistors and capacitors or
resistors and inductors, which we shall refer to for brevity as RC or RL circuits,
respectively. The application of Kirchhoff's laws to these networks gives rise to
differential equations that, in general, are more difficult to solve than the algebraic
equations encountered in the previous chapters. Several methods of solving these
equations will be presented.
We shall first concern ourselves with source-free RC and RL circuits, so-called
circuits in which the forcing or driving functions are constant independent sources
that are suddenly applied to the networks. We shall find that the responses in
Y. these networks consist of two parts, a natural response, similar in form to that of
the source-free case, and a forced response, which is characterized by the forcing
function.
We shall begin our study of a source-free network by considering the series connection
of a capacitor and a resistor, as shown in Fig. 8.1. In this circuit, the capacitor is
charged to a voltage of V at time t = t . Since there are no current or voltage sources
in the network, the circuit response is due entirely to the energy which is stored in the
160
Chap. 8 Simple RC and RL Circuits 1 61
Let us now determine v(t) and /(/) for / > t . Applying KCL at the top node, we
find that
or
% + mv = ° < 8 -2 >
*- ~
_L V
dt RC
from which we obtain
dv
We can now obtain the indefinite integral of both sides of this equation, given by
d
\ i=-TCc\ d, + K <8 4> '
For the solution to be valid for / > / , we see that K must be selected such that the
initial condition of v(t Q ) = V is satisfied. Therefore, at t= we have
f ,
K= '« v° +
m
1
In v — In Vn = In —= -
t -t
RC
If we now recall the relationship
it is apparent that
(8.5)
In Fig. 8.1, we see that this is the voltage across R; therefore the current is
Vo
e -u-
to) IRC
R
L--icl* (8.6)
where the integrals in this equation are definite integrals. Performing the integrations,
we have
In v lnV =
RC
which is equivalent to (8.5).
A graph of (8.5) for t = is shown in Fig. 8.2. We see that the voltage is initially
VQ and that it decays exponentially toward zero as / becomes large. The rate at which
the voltage decays is determined solely by the product of the resistance and the
is characterized by the circuit elements
capacitance of the network. Since the response
and not by an external voltage or current source, the response is called the natural
response of the circuit.
i(t)
v(t)
FIGURE 8.1 Source-free RC circuit FIGURE 8.2 Graph of the voltage response for
fo = in the simple RC circuit of Fig. 8.
Chap. 8 Simple RC and RL Circuits 163
Wj,(oo) = f
p R (t) dt
= -$CVl£
icvi r
which is, indeed, equal to the energy initially stored in the network.
-K~
EXERCISES
In Fig. 8.1, Iet/o = V = 10 V, R = kQ, and C= Find (a) w(10- 3 ),
8.1.1
W s
).
1
8.1.2 The capacitor is charged to a voltage of 100 V prior to the closing of the switch.
For t > 0, find (a) v(t), (b) /(/), (c)wc (t), and (d) the time at which v(t) = 50 V.
Ans. (a) lOOe" 100 ' V, (b) -O.Ole" 100 A, '
(c) 5e- 200! mJ, (d) 6.93 ms
1 =
10k
EXERCISE 8.1.2
8.1.3 The switch in the network opens at / = 1. For t> 1, find (a) t>(l
+
), (b) v(t),
and (c) w c (t). (d) Sketch i(t).
Ans. (a) 10 V, (b) lfor"" 1 '
V, (c) 0.5e- 2 "-" J
164 Simple RC and RL Circuits Chap. 8
10 v — 90 n
EXERCISE 8.1.3
/(/) = ^R e
-' /RC
v(t)
fcSs^ ^ RC= k
V\^><^ RC 2k =
\y ^^\?^\^ / RC = 3K
0.368 Vr
12
i
i —-
i i
3 t/k
Clearly, the current decreases in the same manner as the voltage. It should be noticed
that changing R and C such that the product of RC remains constant causes a change
in the initial current VJR. The current response, however, still decreases in the same
fashion because e~' RC is unchanged.
The time required for the natural response to decay by a factor of \/e is defined as
the time constant of a circuit, which we shall denote by t. In our case, this requires that
which yields
x = RC
The units of r are fi-F = (V/A)(C/V) = C/A = s\ In terms of the time constant, the
voltage response is
Go
K
v{f) V e-'^ (8.7)
The response at the end of one time constant is reduced to e' — 0.368 of its 1
value. Therefore, after four or five time constants, the response is essentially zero.
An interesting property of exponential functions is shown in Fig. 8.4. A tangent to
the curve at t = intersects the time axis at / = t. This is easily verified by considering
the equation of a straight line tangent to the curve at t = 0, given by
i>j = mt + V
where m is the slope of the line. Differentiating v{t), we have
dv
= _YjL e -th
dt x
Therefore
dv
m=
di
and
__ ^_o
T
From Fig. 8.4, we see that an alternative definition for the time constant is the time
required for the natural response to become zero if it decreases at a constant rate
equal to the initial rate of decay.
As a second example, let us consider the more general RC circuit of Fig. 8.5(a).
Prior to the opening of the switch, the circuit is in a dc steady-state condition, by
which we mean that the currents and voltages are constant. As we shall see in Sec. 8.4
given by
eq ^3+ 10Q
(2 + 4)
isn t = o
AAAr/^ 1
IOOV
%J
(a)
-°-
=z ioov
(b)
Since the capacitor is an open circuit to steady-state dc, we see in Fig. 8.5(b) that at
/ =
0" the capacitor voltage (by voltage division between R cq and the 15-fi resistor)
is simply
* ->=(totb) x100=40V
y = V (0 + ) = W (0-) = 40 V
For t > the time constant for the network is simply the product of the capaci-
tance and the equivalent resistance, given by
t =R C= eq 10 s
EXERCISES
8.2.1 In a series RC circuit, determine (a) T for R = 10 kfi and C= 10 //F, (b) C
for R= 100 kft and x = 10 ^s, and (c) R for v(t) to halve every 10 ms on a
1-jxF capacitor. Ans. (a) 0.1 s, (b) 100 pF, (c) 14.43 kfi
8.2.3 The circuit is in a dc steady-state condition at t = 0". For / > 0, find (a)
+
v{0 ), (b) the time constant z, and (c) v,(?).
IOO v^
EXERCISE 8.2.3
1 68 Simple RC and RL Circuits Chap. 8
In this section we shall study the series connection of an inductor and a resistor, as
shown in Fig. 8.6. The inductor, in this case, is carrying a current I at time / = t . As
in the case of the source-free RC circuit,
no current or voltage sources in the
there are
network, and the current and voltage responses are due entirely to the energy stored in
the inductor. The stored energy at / = t is given by
w L (t ) = \L1\ J (8.8)
L^- + Ri =
dt
or
f + |. = (8.9)
This equation is of the same form as that of (8.2) for the RC circuit. We may
therefore solve by separating the variables. Instead, however, let us introduce a
it
second, very powerful method, which we shall generalize in the next chapter. The
method consists of assuming or guessing (a perfectly legitimate mathematical tech-
nique) a general form of the solution based on an inspection of the equation to be
solved. In guessing a solution, we shall include several unknown constants and 4
determine their values so that our assumed solution satisfies the differential equation
and the initial conditions for the network.
A close inspection of (8.9) reveals that i must be a function that does not change its
form upon differentiation; that is, di/dt is a multiple of i. The only function which
i{t) = Ae st
(8.10)
where A and s are constants to be determined. Substituting the solution into (8.9), we
obtain
{' + %)*" o
r
Chap. 8 Simple RC and RL Circuits 169
The constant A can now be determined from the initial condition i(t ) = I . This
condition requires that
i(t) = I e-«'-"
)/L
(8.11)
i(t) = 7 e-"-'°
)/r
in R, in contrast to the RC circuit, lowers the value of the time constant. A graph of a
Therefore the energy absorbed by the resistor as time becomes infinite is given by
u-(oo) = )
p(t) dt
= -^Hq J
Comparing this result to (8.8), we see that the energy initially stored in the inductor is
i(t)
368 I
v(t)
Suppose we had chosen to find the inductor voltage v in the circuit instead of the
v 1 r
,
dt + 7(0) =
which is an integral equation. Differentiating this equation with respect to time, we see
that
1 dv 1 A
+
.
-Rd7 Tv = °
or
This equation is a differential equation that we can solve using one of the methods
discussed previously. It is also interesting to note that if we replace v by iR, we have
di R = r\
Tt
+,
T l
j(0 )
+ = /'(()-) =2A
The time constant for the network for t > is clearly the ratio of the inductance
and the equivalent resistance as seen from the terminals of the inductor. The equivalent
resistance is
R eq = 50 + ^lf =
(
100ft
T = jp = 0.1s
Therefore, since I = +
z(0 )
t
= 2 A and t = 0, we have
Summing the voltages of the inductor and the 50-Q resistor, the voltage v(t) is given by
V (t)= 10 j- + 50/
= -100e- 10 'K
As a final example, consider the network of Fig. 8.9, which contains a dependent
voltage source. The initial current is z'(0) = I . Summing the voltages around the loop,
we find that
L~
di
dt
+
^ Ri + ki =
or
£+m<=°
IOOV
Comparing this equation to (8.9), we see that the equations are identical if R in (8.9)
is replaced by R— k. Thus, from (8.11) with t = 0, we have
i(t) = I e~
{R+k),/L
The time constant in this case, which is modified by the presence of the dependent
source, is given by
= L
x
R+ k
EXERCISES
8.3.1 In a series RL circuit, determine (a) T for R = 200 D. and L = 10 mH, (b) L
for R = 1 kQ and x = 100 ms, and (c) R for the stored energy in a 0.1 -H induc-
tor to halve every 10 ms. Ans. (a) 50 /is, (b) 100 H, (c) 3.47 ft
n
desired to halve the inductor voltage without changing the current response.
Find the new values of inductance and resistance required.
Arts. 0.5 H, 500 Q.
-AW
200 n
i(t) =
IH
~ m 7^—
t
— f
(t) £200
£200. :20on loon iooo 6A
EXERCISE 8.3.3
8.3.4 The circuit shown is in a dc steady-state condition at / = 0~. Find v(t) for t > 0.
^)< t =
ion
0.5v(f)
^ 10 V
EXERCISE 8.3.4
In the preceding sections we have considered source-free circuits whose responses have
been the result of initial energies stored in capacitors and inductors. All independent
current or voltage sources were removed or switched out of the circuits prior to
finding the natural responses. It was shown that these responses, when arising in
circuits containing a single capacitor or inductor and an equivalent resistor, die out
with increasing time.
In this section we
shall examine circuits which, in addition to having initial stored
energies, are driven by constant independent current or voltage sources, or forcing
functions. For these circuits we shall obtain solutions which are the result of inserting
or switching sources into the networks. We shall find that the responses in these cases,
f'
Chap. 8 Simple RC and RL Circuits 173
unlike those of source-free circuits, consist of two parts, one of which is always a
Let us begin by considering the circuit of Fig. 8.10. The network consists of the
parallel connection of a constant current source and a resistor which is switched at
time ! across a capacitor having a voltage v(0~) = V V. For / > 0, a nodal
equation at the upper node is given by
V
c £+ = /o
~R
or
dv ,
I
_ h
dl+RC V C
(8.12)
t=0
X
I ..,
'•I +
R c~ ^ v(t)
) \
;
Equations of this type that have constant forcing functions (7 in this case) can be
solved by the method of separation of variables. We may first write (8.12) in the form
dv Rh
di RC
Multiplying both sides by dt/(v — RI ) and forming indefinite integrals, we have
ln(i'-i?/ )= ~ +K
This result can be written as
v - RI = e
(l IRC) + K
v = Ae~ Rh (8.13)
two parts, an exponential function and a constant function. The exponential function
is of the identical form as that of the natural response in a source-free circuit composed
of R and C. Since this part of the solution is characterized entirely by the RC time
constant, we shall refer to it as the natural response v„ of the driven circuit. As in the
case of the source-free circuit, this response approaches zero as time increases.
The second part of the solution, given by RI , bears a close resemblance to the
forcing function I . In fact, as time increases, the natural response disappears, and the
solution is simply RIQ . This component is due entirely to the forcing function, and we
shall call it the forced response v f of the driven circuit.
Let us now evaluate the constant A in (8.13). As in the case of the source-free
circuit, its value must be selected so that the initial voltage in the circuit is satisfied. At
t = +
, we see that
v(0+) = v(0~) = V
Therefore, at t = +
, (8.13) requires that
V =.A + RI
or
A = V - RI
Graphs of v n v f and v are shown in Figs. 8.11(a) and (b). In (a) the natural
, ,
response v n for V — RI > and the forced response v f are shown. In (b) the com-
plete response is shown.
The current in the capacitor for t > is
M) R*0
r -t/RC
dt R
he resistor is
l
R
= *0 l C = /o
_|_ *<) ^"0 „-tlRC
It is interesting to note that the resistor voltage has changed abruptly from RI at
/ = 0" to V at t = +
. The capacitor voltage, as previously pointed out, is continuous.
The solutions that we have encountered so far in this chapter are often referred to
in other more descriptive terms. Two such terms that are very popular are the transient
response and the steady-state response. The transient response is the transitory portion
of the complete response which approaches zero as time increases. The steady-state
response, on the other hand, is that part of the complete response which remains after
the transient response has become zero.
Chap. 8 Simple RC and RL Circuits 1 75
n "' '«
v -Rl
(a)
Vo
RI,
(b)
In our example, we see that the transient response and the natural response are
identical, as are those for the source-free circuits of the previous sections. The steady-
state response is therefore identical to the forced response. In our example this
response is RI , a dc value which we defined as a dc steady-state condition in Sec. 8.2.
We should not conclude from the above discussion that the natural and forced
responses are always the transient and steady-state responses. If the forcing function
is a transitory function, for instance, the steady-state response is zero, as we shall see
in the next chapter. In this case the complete response is the transient response.
EXERCISES
8.4.1 In Fig. 8.10 I = 100 mA and R = 1 kQ. For t > 0, determine (a) V such
that the natural response is zero, (b) /c (10
-4
) for V = 50 V and C= 1 /zF,
and (c) C for VQ = -50 V such that v(W~ 3 ) = 0.
8.4.2 Find v(t) for t > if v(0~) = 5 V. Am. 10 - 5e- 100 'V
t =
ioi<n
D l/f i A-
EXERCISE 8.4,2
The equations describing the networks of the previous sections are all special cases of a
general expression given by
(8.15)
instance, we compare this equation to that of (8.2) for the source-free circuit of Sec.
8.1, we see that y = v, P = /RC, and Q = 0. The same relations are valid for the
1
We see that if we multiply both sides of (8.15) by e Pl the left-hand side becomes a ,
(ye"') = Qe>
dt
ye f = Qe p '
dt
J
Chap. 8 Simple RC and RL Circuits 177
y = e
-p
'
I"
Qe p '
dt + Ae~ (8.16)
(8.17)
/\*s\ ft
where y„ =
Ae~ Pt and y f = Q/P are the natural and forced responses. We observe
tKaTTn has the same mathematical torm as the source-free natural response and that
y f is always a constant which is proportional to Q. In addition, \jP is the time constant
in the natural response.
To illustrate the method, let us find i2 for t > in the circuit of Fig. 8.12, given that
':(0) = 1 A. Although the somewhat complex combination of elements, the
circuit is a
solution of (8.17) is valid since the network contains a constant forcing function and a
single energy-storage element (the inductor). The loop equations for the circuit are
8i, - 4/ 2 = 10
a\
-4/, + 12/j =
dt
Eliminating i t
from these equations, we find that
di 2
+ 10/ 2 = 5
dt
Comparing this equation with (8.15), we see that P= 10 and Q= 5. Hence (8.17)
yields
i2 = Ae~ 10 '
+ £
i
a (0)
= A + $=1
1 78 Simple RC and RL Circuits Chap. 8
ii = y- io,
+
EXERCISES
8.5.1 Find v(t) for t > if v(0) = 0. Ans. 12(1 -<?-"") V
EXERCISE 8.5.1
8.5.2 Find iz (t) in Fig. 8.12 if the source voltage is lOe"' V and / 2 (0) = 1 A.
Ans. \e-< + %e~ 10 '
A
8.5.3 Find /(?) for / > if v(0) = 12 V. Ans. 2 - <r 5 < A
3A
EXERCISE 8.5.3
Let us now introduce a shortcut procedure that is very useful for finding the currents
and voltages in many circuits. The technique involves formulating the solution by
merely inspecting the circuit. Consider, for instance, the example of the previous
section (Fig. 8.12). We know that
k = hn + Hf
where i 2n and i
2f are the natural and forced responses, respectively. Since i 2n has the
Chap. 8 Simple RC and RL Circuits 179
same form as the source-free response, we can look at the network in the absence of the
forcing function (i.e., make the 10-V source zero by replacing it by a short circuit), as
shown in Fig. 8.13(a). The natural response is then
'In
concerned, it does not matter at what time we look at the circuit. We may choose then
to look at the circuit in the steady state when i 2 „ is zero. At this time the inductor is a
short circuit, as shown in Fig. 8.13(b), and
'2/ \
Therefore
i2 = Ae- i0 '
+ £
The constant A is now determined as before from the initial condition, z'
2 (0)
= 1.
(a) (b)
FIGURE 8.13 Circuits for finding the response of Fig. 8.12. (a) Circuit
for finding ij. n ; (b) circuit for finding iif
As a second example, let us find i{t) for t > in Fig. 8.14, given v(0) = 10 V. The
current is given by
* = K +h
60 a
v(t)
To obtain /„ we note that it has the same form as v„, the natural response of the
capacitor voltage. In fact, the natural response of every current or voltage in the
circuit has the same form as v„. This is true because all the other currents and voltages
in the source-free circuit may be obtained from v„ by applying one or more of the
operations of addition, subtraction (in KCL and KVL), differentiation, and integra-
tion, none of which changes the nature of the exponential e~"\ Examining the source-
free circuit (the current source open circuited), we see that the time constant for the
capacitor voltage is x — 100 s. Therefore
= Ae -(/100
In the steady state the capacitor is an open circuit, and the forced response is, by
inspection,
// = A 1
Therefore
To evaluate A, we must find the value of /(0 + ). Since v(0) = v(0 + ) = 10 V, summing
the voltages around the right-hand mesh, for t = +
, we have
/(0
+
) = 0.7
0.7 = 1 + A
Therefore A = —0.3 and
/(/)= 1 -0.3<?-'/10 °
Before concluding this section let us, as a final example, determine /'(/) and v{t) in
/(0-) =2A
and
v(0~) = 60 V
Chap. 8 Simple RC and RL Circuits 181
When the switch closes at / = 0, we observe that nodes a and b are short-circuited
together, and we can redraw the network as shown in Fig. 8.15(b). It should be
noticed that the 30-fl resistor need not be included in this circuit because the switch is
a short circuit across its terminals. The combination is equivalent to a 30-fi resistor in
parallel with a 0-Q resistor, which, of course, is Q or a short circuit.
io n
-/VW-^
f-
I H 3on:
*A
t = 3F v(t) 2on
© bA
(o)
(b)
Let us next consider the current i(t) leaving a in Fig. 8.15(b) through the 15-Q
resistor. From KCL, this same current must enter a through the 1-H inductor; hence
no current flowing in the circuit to the left of a can enter the other part of the circuit to
the right of o, and vice versa. Thus after the switch is closed, the network reduces to
two independent circuits, each of which can be solved individually.
The first circuit, consisting of the 1-H inductor and the 15-Q resistor, is simply a
source-free RL network having /(0 ) = z'(0") = 2A. Therefore
+
The second circuit, composed of all the elements to the right of a, is simply a driven
RC network with r(0 + ) = v(Q~) — 60 V. From our shortcut procedure, we find
EXERCISES
8.6.1 Find v(t) for / > if the circuit is in dc steady state at / = 0~.
5
Arts. 2.5(1 -e-'° ')V
10V 10mA
EXERCISE 8.6.1
8.6.2 Find /(/) and v(t) for / > if the circuit is in dc steady state at t = 0~.
Aaa,
lOOkO
i(t)
r t
lOOmH
nnnp
v(t)
— »
l.5kn
vw-
EXERCISE 8.6.2
In the previous sections we have analyzed circuits in which energy sources have been
suddenly inserted into the networks. At the instant these sources are applied the
voltages or currents, at the points of application, change abruptly. Forcing functions
whose values change in this manner are called singularity functions. There are many
singularity functions that are useful in circuit analysis. One of the most important is
the unit step function, so named by the English engineer Oliver Heaviside (1850-1925).
The unit step function is a dimensionless function which is equal to zero for all
negative values of its argument and which is equal to one for all positive values of its
argument. If we denote the unit step function by the symbol u{t), then a mathematical
description is
u(t) = 0, t<
(8.18)
= 1, t>0
Chap. 8 Simple RC and RL Circuits 183
From a graph of (8.18), shown in Fig. 8.16, we see that at t = 0, u{t) changes abruptly
from to 1. Some authors define u(0) to be 1, but we are leaving w(/) undefined at
( = 0.
u(t)
terminals. We have assumed in our model that the switching action occurs in zero
time.
o t =o
J<
Vu(t)(+) V -^
Vu(l)
(a) (b)
Equivalent circuits for a current step source of /amperes are shown in Fig. 8.18.
An open circuit exists for / < 0, and the current is zero. For t > 0, the switching
action causes a terminal current of / amperes to flow.
Iu(t)
t=o
<
;u(,)
(t)
©
(a) (b)
The switching action shown in Fig. 8.17 can only be approximated in actual
circuits. However, in many cases, it is not necessary to require that the voltage source
be a short circuit for / < 0, as we shall see in the next section. If the terminals of a
network to which the source is to be connected remain at V for / < 0, then a series
connection of a source V and a switch is equivalent to the voltage step generator, as
shown in Fig. 8.19.
t =
r +
v(t)
Network
(v(t)=0 Vu(t)( Network
for t<0)
(a) (b)
Equivalent circuits for a current step generator in a network are shown in Fig.
.20. In each case the current in the network terminals must be zero for / < 0.
Network
w-o (
for
i(t) =
t <0)
Iu(t)(f Network
(o) (b)
Let us now return to our definition of the unit step function given in (8.18). We may
generalize this definition by replacing t by t — t in the three places that it occurs,
which results in
The function u(t — t Q ) is the function u(t) delayed by t seconds, as shown in Fig. 8.21.
Multiplying (8.19) by Kor /gives us a voltage step source or a current step source
whose value changes abruptly at time t . Equivalent networks for these sources are
obtained in Figs. 8.17-8.20 by taking all actions related to switching to occur at
t = t .
Chap. 8 Simple RC and RL Circuits 1 85
u(t-t n )
Step functions are very useful in formulating more complex functions. Take, for
instance, the rectangular voltage pulse of Fig. 8.22(a). From this figure, we see that
/ <
= V, < < t t
= 0, t > t
Since u(t) becomes 1 for / > and — u(t — t ) becomes —1 for t > t , it is obvious
that
Vl (t) = V(0 - 0) =
for < < t t ,
Vl (t) =V(l-0)=V
and for / > t ,
v t (f) = V{\ - 1) =
Now suppose that we wish to produce a train of these pulses with one occurring
every T seconds, where T> / , as shown in Fig. 8.22(b). Such a wave is called a
square wave. The first pulse is given by (8.20). The second pulse is simply the first pulse
delayed by T seconds. Therefore replacing t by t — T \n (8.20). we have
The (n ~ 1 )th pulse in the pulse train is the first delayed by nT, and therefore
To obtain an expression for the square wave for all / > 0, we add the above expres-
sions and obtain
v.(t)
(a)
v (t)
2
r
1
K T T+ t 2T 2T + t„
(b)
Waveforms like those of (8.20) and (8.21) are very common in digital circuits such
as those in the digital computer.
EXERCISES
8.7.1 Using unit step functions, write an expression for the current i'(f) which satisfies
= 0, / > 1 s
8.7.3 Using unit step functions, write an expression for v(t) for — oo < < oo.
t
10 sin 2nt
The step response is the response of a circuit having only one input which is a unit
step function. The response and step input can, of course, be a current or a voltage.
The is due entirely to the step input since no initial energies are present
step response
in the dynamic circuit elements. This is the case because all the currents and voltages
in the network are zero at / = 0~ due to the fact that the step function is zero for
— oo < / < 0. In this section we shall consider step responses for which the step input
may or may not be a unit step.
As an example of a step response, let us find v(t) in the simple RC circuit of Fig.
8.23(a) having a voltage step input of Vu{t) volts. Applying KCL, we have
r dv v - Vu(t)
_ n
C
dt
+ "
R ~
U
or
dV V V
+ RC = RC u{t)
.
f.y.
dl
v{t) = 0, t <
188 Simple RC and RL Circuits Chap. 8
which confirms our assertion that the response is zero prior to the change in the
input.
For t > 0, our differential equation is
dv V
dt
r RC ~ RC
We know that
V = vn + v,
where
v, ,
= Ae~'/RC
and, by inspection,
vf = V
Therefore
v= V+ Ae' t/RC
The initial condition v(0 + ) = v(0~) = requires that A = —V, and therefore our
solution for all / is
v(t) = 0, t<0
= V[\ -t/RC
I t>0
This may be written more concisely, using the unit step function, as
The voltage across the resistor and the capacitor is zero for t < 0. Therefore an
equivalent circuit for our network issatisfied by the circuit of Fig. 8.23(b) provided we
specify that v(Q~) — 0.
(a) (b)
FIGURE 8.23 (a) RC circuit with a voltage step input; (h) equivalent circuit
As a second example, let us find v 2 {t) in the circuit of Fig. 8.24, consisting of a
resistor, a capacitor, and an op amp. A nodal equation at the inverting terminal of the
Chap. 8 Simple RC and RL Circuits 1 89
op amp is given by
7f +C f= °
since the node voltage and the current of the inverting terminal are both zero.
Therefore
V_
v 2 (0
RC
u(t) dt +K
V_
RC
tu(t) +K
=-
v2
m tu(t)
A graph of v 2 is
shown in Fig. 8.25. This function is called a ramp function with a slope
of - Vj RC.
As a final example, let us find the voltage v{t) in the network of Fig. 8.26, given that
190 Simple RC and RL Circuits Chap. 8
2H< v(t)
j'(0") = 0. The forcing function for the circuit is the current pulse,
i
g {t)= \0[u(t)-u(t- 1)]A
zero, and the response is simply the source-free response resulting from the energy
which the inductor has accumulated during the interval of the current pulse. There-
fore we see that the solution of the problem involves finding the step response of the
circuit for / < 1 s and then finding the source-free response for t > 1 s.
We know that the step response is of the form \
v = v„
where
v„ = Ae-*">'
/L
= Ae~ s " s Ae
„ _ 2
r (3)(io) n = 12
Chap. 8 Simple RC and RL Circuits 1 91
v = Ae~' + 12
v = 0, / <
= 12(1 - e-'), < < / 1
t; = Ar'
=
\
At t 1". our solution for the step response gives
v(\
+
) = Be' 1
= 12(1 - <?">)
or
B= 12(1 -*-»)*
(t)
IOA
cuit to i
g (t)
1 92 Simple RC and RL Circuits Chap. 8
v= 12(1 - r'K 1
'- 11
, t > 1
EXERCISES
8.8.1 Find the step response i{t) to the voltage step v(t) = 20u(t) V.
Am. (1 -0.5£>- 100 'M0mA
</0) «,(t)
EXERCISE 8.8.1
8.8.2 Find the step response v 2 {t) to the voltage step v t (t) = I0u(t) V. Sketch v 2 (t)
for < t < 1(T 2 s. Ans. (10 + 100f)«(/)
i0kn
EXERCISE 8.8.2
8.8.3 Find v x
{t) in Ex. 8.8.1 if v(t) =
20[u(t) - u(t - 0.1)] V. Sketch Vl (t).
Ans. 10[1 -
e- i00 '][u(0 «(' ~ - 0.1)] + 10(e 10 - l)e- 100 'u(t - 0.1) V
Chap. 8 Simple RC and RL Circuits 193
In this section we shall consider the use of superposition for obtaining solutions of RC
and RL circuits containing two or more independent sources. As a first example, let us
consider the circuit of Fig. 8.26 of the previous section. The value of the independent
current source is given by
i, = I0u(t) - \0u(t - l)
', = 'i + h
where ;', = lCh/(/) and i2 = -\0u(t — 1), the circuit of Fig. 8.26 can be redrawn as
shown in Fig. 8.28.
v = v x
+ v2
where i^ and v 2 are the responses due to /, and 2 respectively. In our previous z ,
solu-
tion we found that the step response due to the current step /, was given by
v, = 12(1 - e-')u(t)
Next, we need the response v 2 due to i2 . We note that i 2 is simply the negative of /,
v2 = -12(1 - e- u -")u{t - 1)
As a second example, let us consider the RC network of Fig. 8.29, which contains
two independent sources and an initial capacitor voltage v(0) — V Applying KVL .
(R, +R 2 )i +
l
-^ idt+V =V 1
-R 2IX
Clearly, the current response becomes Ki when the independent sources and the
initial capacitor voltage are multiplied by the factor K, which demonstrates the
proportionality property for a linear network. This result is easily extended to any
linear circuit containing one or more capacitors. Thus initial capacitor voltages can be
treated as independent voltage sources. In a similar manner, it is easily shown that
initial inductor currents can be treated as independent current sources.
We shall now employ superposition to determine the voltage v by finding v u v 2 ,
and v 2 the responses due to V ,I U and V respectively. The circuit for finding v is
,
X
,
1
shown in Fig. 8.30(a). This is a simple driven RC circuit having a zero initial capacitor
voltage. The solution is given by
Vl = K,(l - e
- ,nRl+R ' )c
)
The circuit for finding v 2 is shown in Fig. 8.30(b). This again is a simple driven
circuit having a zero initial capacitor voltage, for which we find
v2 = -* 2 /iO - e
-t/{R + X i )C
l
In the circuit of Fig. 8.30(c), the voltage v z is simply the source-free response
resulting from the initial capacitor voltage. Since v 2 (0) = V , we find that
— V
V e -tHRi + Ri)C
Chap. 8 Simple RC and RL Circuits 195
FIGURE 8.30 Circuits for finding (a) V\, (b) v 2 . and (c) v 3 for an RC network
V = +v +v
i>, 2 3
= v - rj, +
x (/? 2 /, - v + r )*-'
x
/0, ' +JWC
(8.24)
vf = v lf + v 2f
and
Vif = -RJx
Therefore the forced response is
vf = V, - i? 2 7,
The natural response, obtained from the source-free circuit [Fig. 8.30(c)], is given
by
y — U+Rt)C
^g-'/(i!l+«2)C
Therefore
v=V - x
R 2 I, + Ae-'/(* l+Rt)c (8.25)
EXERCISES
8.9.1 Reduce (8.22) to the form of (8.23).
8.9.3 Use superposition to find i for t > 0. Assume the circuit is in a steady-state
condition at / = 0" Am. 5 + 10(1 - r s6o ')mA
P
Vji-WU 1st
EXERCISE 8.9.3
%A> H*
i ^ 7*
•>
*5^ rVt
,
PROBLEMS
»> -H-xi? 1*5
ft
-2
.1 The current i(10 ) = 3.68 mA. Determine (a) t<0), (b) v(r) for f > 0, and (c)
the energy dissipated by the 10-kQ resistor as t becomes large, (d) Sketch v(t)
forr>0.
ft!
t = o
v(t) io kn
i
PROBLEM 8.1
2
8.2 Determine the new value of resistance necessary in Prob. 8.1 for r(10 )
~
\
\A
Chap. 8 Simple RC and RL Circuits 1 97
PROBLEM 8.3
8.5 The circuit is in a steady-state condition at / = 0". Find i(t) for / > 0.
PROBLEM 8.5
8.6 Consider a source-free circuit that has a response v(t) = V e~' /r . Show that a
straight line that is tangent to a graph of this response at time t {
intersects the
time axis (abscissa) at time /] + T.
8.7 The circuit is in a steady-state condition at / = 0". Find /'(/) for / > 0.
i(t)
1 2 .ft 2H
OlfP-
n
V
PROBLEM 8.7
'V
8.8 Find v(t) for / > 0, given /(0) = 2 A.
-
i(t)
3ft
Aa/v 1
v(t)
-o
4H3 v(t)
PROBLEM 8.8
CO:) --
f\
;
1 98 Simple RC and RL Circuits Chap. 8
i(t)
O.I H 2i(t)A
PROBLEM 8.9
xl0 i(t)V
3 y v(t)
PROBLEM 8.10
\AA
PROBLEM 8.11
8.12 Repeat Prob. 8.11 with the 0.25-F capacitor replaced by a ^-H inductor. Find the
inductor current.
8.13 Repeat Prob. 8.11 with the 6-V source replaced by 6e~' V.
lOi.V
4A
PROBLEM 8.14
' *
3 1H /YO -
&?> ' V, f "L-rt-
~"^A— —
Chap. 8 i Simplte RC and RL Circuits 199
8.15 Repeat Prob. 8.14 with the 4-A source replaced by 6e~' A.
V.^ I mH 5 n 5 n
(Jjp
(l) 20v
i/ - £•.
'^^
PROBLEM 8.16 ^< +
8.17 Express r(/) in terms of unit step functions, where ^
= ©
'
P -
-
£ ft) ~ov, /<-io iic'-fZ^-rf'
10 V, 10 < <0 /
2-
= f
W K =
20V,
15V,
< < 10
10 <'
t
6U.
8.18 Express i(t) in terms of unit step functions.
«(0 = S
n=0
(' - «)[«(' - « - 1) -u{t-n - 2)] ^
8.20 Find i for t > if v = 24«(r) V.
i2n
PROBLEM 8.20
W
200 S/>77p/e RC and RL Circuits Chap. 8
Ca ' '
_u(t)(p
r' ^ t/<LC^
PROBLEM 8.22
i/
c 8.23 Using superposition, find v for t > 0, if v(0) = 0.
V .* + * '
-v^
U' 8.24 Find u for ? > 0. Assume the circuit is in a steady-state condition at t = 0"
>/(.oY
vu t =
2kH IKfl
\&
10 V IOmA
f^
1*
y% PROBLEM 8.24
* *
U>
9
SECOND-ORDER CIRCUITS
In the case of linear circuits with energy-storage elements the describing equations
(those relating the outputs to the inputs) may be expressed as linear differential
equations, because the terminal relations of the elements are such that the terms in
the loop or nodal equations are derivatives, integrals, or multiples of the unknowns
and the source variables. Evidently a single differentiation of an equation will remove
any integrals that it may contain, so that in general the loop or nodal equations for a
given circuit may be considered to be differential equations. The describing equation
then may be obtained from these equations.
The circuits containing storage elements that we have considered so far were
first-order circuits. That is, they were described by first-order differential equations.
This is always the case when there is only one storage element present or when a
switching action converts the circuit into two or more independent circuits each having
no more than one storage element.
In this chapter we shall consider second-order circuits, which as we shall see,
contain two storage elements and have describing equations that are second-order
differential equations. In general, //th-order circuits, containing n storage elements,
are described by mh-order differential equations. The results for first- and second-
order circuits (n =
1 and n — 2) may be readily extended to the general case, but we
To introduce the subject of second-order circuits, let us begin with the circuit of Fig.
9.1, where the output to be found is the mesh current z*
2 . The circuit contains two
201
202 Second-Order Circuits Chap. 9
storage elements, the inductors, and as we shall see, i 2 satisfies a second-order differ-
ential equation. Methods of solving such equations will be considered in later sections
of this chapter.
The mesh equations of Fig. 9.1 are given by
2^+13,-4/, = *,
, (9.1)
-4/, + % + 4/ = 2
4% + 4^
dt
" 4 V dt
2 (9.3)
Substituting (9.2) and (9.3) into the first of (9.1) to eliminate /,, we have, after multi-
plying the resulting equation through by 2,
The describing equation for the output i2 is thus a second-order differential equation.
That is, it is a differential equation in which the highest derivative is second-order.
For this reason we shall refer to Fig. 9.1 as a second-order circuit and note that,
typically, second-order circuits contain two storage elements.
There are exceptions, however, to the rule that two-storage-element circuits have
second-order describing equations. For example, let us consider the circuit of Fig.
9.2, which has two capacitors. With the reference node taken as indicated, nodal
equations at the nodes labeled v and v 2 are given by
x
£ + .,«.. (9.5)
$h. +
_i_ 2v 2 -2v *
dt
The choice of thenode voltages v and v 2 as the unknowns has resulted in two
x
FIG U R E 9.1 Circuit with two inductors FIGURE 9.2 Circuit with two capacitors
Evidently Fig. 9.2, although it contains two storage elements, is not a second-
order The same voltage v g is across each RC combination, and thus the circuit
circuit.
may be redrawn as two first-order circuits. If the source were a practical source rather
than an ideal source, then the circuit would be a second-order circuit. (See Prob. 9.5.)
EXERCISES
9.1.1 Find the equation satisfied by the mesh current i2 .
dh\ dvg
Ans.
dt 2
+ 7f + 6;2
dt
EXERCISE 9.1.1
In Chapter 8 we considered first-order circuits in some detail and saw that their describ-
ing equations were first-order differential equations of the form
204 Second-Order Circuits Chap. 9
4* + a x=f(t) (9.6/
In Sec. 9.1 we denned second-order circuits as those having two storage elements with
describing equations that were secon d-order differential equa tions, given generally by
yA In -fU^ -ferr^
f^g + ^f + Oox=f(t)
} (9.7)
In (9.6) and (9.7) the a's are real constants, x may be either a voltage or a current, and
f{t) is a known function of the independent sources.
As an example, for the circuit of Fig. 9.1, the describing equation was (9.4).
Comparing this equation with (9.7), we see that a = 10, a Q = 16, fit) = 2vg and x ,
x = i2 .
From Chapter 8 we know that the complete response satisfying (9.6) is given by
x = x„ + xf (9.8)
where x n is the natural response obtained when/(/) = and x f is the forced response,
which satisfies (9.6). The forced response, in contrast to the natural response, contains
no arbitrary constants.
Let us see if this same procedure will apply to the second-order equation (9.7).
By a solution to (9.7) we shall mean a function x which satisfies (9.7) identically. That
is, when x is substituted into (9.7), the left member becomes identically /(/)• We shall
also require that x contain two arbitrary constants since we must be able to satisfy
the two conditions imposed by the initial energy stored in the two storage elements.
If x n is the natural response, i.e., the response when/(/) = 0, then it must satisfy
the equation
d 2 x„ dx
+ a ^t
,
, a xn = (9.9)
~dF
Since each term contains x„ to the same degree, namely 1 (the right member may be
thought of as = 0x n ), this equation sometimes called the homogeneous equation.
is
dt
*f.
2 + ai £?
'
dt
+a '
xf =fQ) (9.10)
Adding (9.9) and (9.10) and rearranging the terms, we may write
The rearrangement is possible, of course, because the equations involved are linear.
Comparing (9.7) and (9.11) we see that (9.8) is our solution, as it was in the first- \
order case. That is, x satisfying (9.7) is made up of two components, a natural response
Chap. 9 Second-Order Circuits 205
.v„ satisfying the homogeneous equation (9.9) and a forced response x f satisfying the
original equation (9.10) or (9.7). As we shall see. the natural response will contain
two arbitrary constants and. as in the first-order case, the forced response will have
no arbitrary constants. We shall consider methods of finding the natural and forced
responses in the next three sections.
Of course, if the driving, or forcing, functions are such that/(/) = in (9.7), then
the forced response is zero, and the solution of the differential equation is simply the
natural response.
EXERCISES
9.2.1 Show that
x t
=A t
e '
and
xz = A 2 e~ 2 '
& + 3$ + *-0
regardless of the values of the constants, A and A 2
x
.
9.2.3 Show that if the right member of the differential equation of Ex. 9.2. 1 is changed
from to 6, then
x = Aie~' + A 2 e~ 2 + '
3
is a solution.
x = x„ + xf
^ + a^ + a x = (9.12)
Evidently the solution x = x„ must be a function which does not change its form when
206 Second-Order Circuits Chap. 9
it is differentiated. That is, the function, its first derivative, and its second derivative
must all have the same form, for otherwise the combination in the left member of the
equation could not become identically zero for all t.
xn = Ae s '
(9.13)
since this is the only function which retains its form when it is repeatedly differentiated.
This is, of course, the same function that worked so well for us in the first-order case
of Chapter 8. Also, as in the first-order case, A and s are constants to be determined.
Substituting (9.13) for x in (9.12), we have
As 2 e s'
+ Asa e"
x
+ Aa e
s'
=
or
Ae s
'(s
2
+ a :
s + a ) =
Since Ae" cannot be zero [for then by (9.13) x„ = 0, and we cannot satisfy any initial
s
2
+ a,s + aQ = (9.14)
Since (9.14) is a quadratic equation, we have not one solution, as in the first-order
case, but two solutions, say s and s 2 given by the quadratic formula as
x ,
— a ± Ja\ — x
4a
(9.15)
S 1 7.
The coefficients A and A 2 are, of course, arbitrary. Either of the two solutions (9.16)
x
will satisfy the homogeneous equation, because substituting either into (9.12) reduces
it to (9.14).
As a matter of fact, because (9.12) is a linear equation, the sum of the solutions
(9.16) is also a solution. That is,
xn = xnl +x n2 (9.17)
is a solution of (9.12). To see this, we have only to substitute the expression for x„
into (9.12). This results in
Chap. 9 Second-Order Circuits 207
d2 d
x
jp( »i + x »i) + °i ^(x + »i *») + fl °(*»i + a »J
/d 2 x„, . dx„, . \ ,
(d 2 x. 2 , (/x n2 i \
-0+0=0
since both x nl and .v n2 satisfy (9.12).
By (9.16) and (9.17) we have
x„ = A<?
J
" + A ze s "
(9.18)
which is a more general solution (unless s, = s2) than either equation of (9.16). In
fact. (9.18) is called the general solution of the homogeneous equation if s t
and s 2 are
distinct (i.e., not equal) roots of the characteristic equation (9.14).
As an example, the homogeneous equation corresponding to (9.4) in the previous
section is given by
^+10^+16/
dt
+ *2 2
= (9.19)
j2 + 10s + 16 =
The roots are s — —2 and s = —8, so that the general solution is given by
i2 = A^- + 2'
A 2 e-*' (9.20)
The reader may verify by direct substitution that (9.20) satisfies (9.19), regardless of
the value of the arbitrary constants.
Because (9.18) is and s 2 are sometimes called
the natural response, the numbers s,
EXERCISES
9.3.1 Given the linear differential equation
d 2x
dt 2
<§-°
show that xi = t
1 and x 2 = 1 are both solutions but that x x + x2 is not a
solution.
9.3.3 Given
<*>©+<§ 3x =
2
d x
(b)
dt
2§ + * =
2
find the characteristic equation and the natural frequencies in each case.
Am. (a) -1, -3, (b) -1, -1
Since the natural frequencies of a second-order circuit are the roots of a quadratic
characteristic equation, they may be real, imaginary, or complex numbers. The nature
of the roots is determined by the discriminant a 2 — 4a of (9.15), which may be posi-
tive (corresponding to real, distinct roots), negative (complex roots), or zero (real,
equal roots).
For example, consider the circuit of Fig. 9.3, where the response to be found is the
mesh current i2 The loop equations, written around the left mesh and the outer loop,
.
are given by
c*+4y,+£-»,-£=..
At, + 4 f i, dt + t<0) = v,
Jo
Substituting for /, from the second equation into the first, we have
1 (dv„ A . \ „. di
(H-V«-\j>*-«> +-H*-*-) 2
dt
Chap. 9 Second-Order Circuits 209
2n
== /n satisfies the homogeneous equation
^ +^
dt
1 K "^ 1)^
(/?+
(/?+l)^f
>
l
dt
+ (/? + 4)/ n =
s2 + (R + \)s +R+ 4 =
Using the quadratic formula, we have the natural frequencies, given by
s ^_ -(R+\)±^R
2
-2R- 15
(92])
s U2 = -2, -5 (9.22)
s U2 = -3, -3 (9.23)
where j = «J—\. (In electrical engineering we cannot use as the mathematicians do,
/',
for the imaginary number unit, since this would result in confusion with the current.
Complex numbers are considered in Appendix C for the reader who needs to review
the subject.)
If the natural frequencies 5, 2 are real and distinct, then the natural response is
in = A.e- 1 '
+ A 2 e~ 5 '
Complex Roots
where a and fi
are real numbers. For example, in the case of (9.24), a =— 1 and fi
=
2. By (9.18) the natural response in the general case is
This appears to be a complex quantity and not a suitable answer for a real current
or voltage. However, because A and A 2
x
are complex numbers, it is mathematically
correct, although somewhat inconvenient.
To put the natural response (9.25) in a better form, let us consider Euler's formula,
given by
e" = cos0+./sin0 (9.26)
These results are derived in Appendix D. They are named for the great Swiss mathe-
matician Leonhard Euler (pronounced "oiler"), who lived from 1707 to 1783. Euler's
greatness is by the fact that the number
attested to e, the base of the natural loga-
rithmic system, was chosen in his honor.
Using (9.26) and (9.27), we may write (9.25) as
x„ = <?«'(iV' + A e- ") z
Jl
Since A and A 2
x
are arbitrary, let us rename the constants as
A = B, x
+ A2
Mi —Mi = B 2
so that
x„ = a
e '(B l
cos fit +B 2
sin fit) (9.28)
in = e~'(B 1
cos It +B 2 sin It)
where B x
and B 2 are, of course, arbitrary.
The last type of natural frequencies we may have are those that are real and equal,
say
j, =s = 2 k (9.29)
Chap. 9 Second-Order Circuits 211
In this case (9. 18) is not the general solution since both x ni and x„ 2 are of the form
Ae kt and
, thus there is only one independent arbitrary constant. For (9.29) to be the
natural frequencies, the characteristic equation must be
(s - k) 2 = s
2 - 2ks + k2 =
and therefore the homogeneous equation must be
^-2k^f+k x 2
n
= (9.30)
xn = h{t)e
k<
#h *kt _
dt 2
Therefore /;(/) must be such that its second derivative is zero for all t. This is true
if /;(/) is a polynomial of degree 1, or
h(t) =A + i
A2t
where A and A 2
l
are arbitrary constants. The general solution in the repeated-root
case, s u2 = k, is thus
xu = {A +A % t)t» l (9.31)
which may be verified by direct substitution into the homogeneous equation (9.30).
As an example, in the case of (9.23) we have s l2 = —3, —3, and thus
iH = (A +A t 2 t)e-
3'
EXERCISES
9.4.1 Find the natural frequencies of a circuit described by
d 2x dx n
Wl +
.
a +.
a x =
x
Tt
if (a) a x
= 5, a = 6; (b) a x = 4, a = 13; and (c) a = 8, a = 16.
x
9.4.2 Find x in Ex. 9.4.1 with the arbitrary constants determined so that x(0) = 1
and dx(0)/dt = 4.
Ans. (a) 7e~ 2 ' - 6e~ 3 ', (b) <r 2 '(cos 3t + 2 sin 30, (c) (1 + 8/)<r 4r
9.4.3 Find x if
d 2x
J+ 9X = °
Ans. x =A i
cos 3/ + A 2 sin 3?
The forced response x f of the general second-order circuit must satisfy (9.10) and
contain no arbitrary constants. There are a number of methods for finding x f but ,
for our purposes we shall use the procedure of guessing the solution, which has worked
so well for us in the past. We know from our experience with first-order circuits that
the forced response has the form of the driving function. A constant source results
in a constant forced response, etc. However, the response must satisfy (9.10) identi-
cally, which means that first and second derivatives of x
f as well as x f itself, will ,
appear in the left member of (9.10). Thus we are led to try as x f a combination of the
right member of (9.10) and its derivatives.
As an example, let us consider the case v g — 16 V in Fig. 9.1. Then by (9.4), for
i 2 = x, we have
^ T+ 10^+16*
dt
2 ^ =
dt
32 (9.32)'
v
Xn = A x
e- 2 ' + A 2 e~" (9.33)
Since the right member of (9.32) is a constant and all its derivatives are constant
(namely zero), let us try
xf =A
where A is a constant to be determined. We note that A is not arbitrary but is a par-
ticular value that hopefully makes x f a solution of (9.32). Substituting xf into (9.32)
yields
16A = 32
or
xf =A= 2
A knowledge of the initial energy stored in the inductors can now be used to evaluate
A and A 2
x
.
In the case of constant forcing functions we may often obtain x f from the circuit
itself. In the example just considered xf is the steady-state value of i 2 in Fig. 9.1 when
vg = 16 V.At steady state the inductors look like short circuits, as shown in Fig.
9.4, so that, from the figure, we have
vg = 20 cos At V
d2 x dx
.
1 + 10^+ 16x = 40 cos 4? (9.34)
dt dt
The natural response x„ is given by (9.33), as before. To find the forced response x f
we need to seek a solution which contains all the terms, and their possible derivatives,
in the right member of (9.34). The coefficients of these terms will then be determined
by requiring x f to satisfy the differential equation. In the case under consideration,
the only term is a cos At term and the trial,
contains this term and all its possible derivatives (which are cos At and sin At).
From (9.35) we have
^= dt
2
-16,4 cos At - 165 sin At
Substituting these values and (9.35) into (9.34) and collecting terms, we have
Since this must be an identity, the coefficients of like terms must be the same on both
sides of the equation. In the case of the cos At terms we have
405 - 40
xf = sin At (9.36)
i2 =x= A x
e- 2 '
+ A 2 e-*' + sin At (9.37)
fit) xf
k A
t At +B
t* At 1 + Bt + C
C" Ae"'
sin bt, cos bt A sin bt + B cos bt
e"' sin bt, e"< cos bt e"'(A sin bt + B cos bt)
EXERCISES
9.5.1 Find the forced response if
£g* + 4 $ + 3,-/<0
where f{t) is given by (a) 6, (b) Ae' 1 ', and (c) 9t.
9.5.2 If x(0) = 2 and dx(0)/dt = 4, find the complete solution in Ex. 9.5.1.
Ans. (a) 2e~' - 2e' 3 '
+ 2, (b) le~< - Ae' 1 - e~ (c) 9.5<r' - 3.5e" + 3/ - 4
' 3
',
3t
Chap. 9 Second-Order Circuits 215
dz x dx
w --(a + b)^ +
,
{a
, ,
b)
v
di
abx=f(t) (9.38)
where a and b =£ a are known constants. In this case the characteristic equation is
s2 - (a + b)s + ab =
from which the natural frequencies are
st = a, s2 = b
x„ = A.e"' + A 2 eb '
(9.39)
where A :
and A 2 are arbitrary.
Let us suppose now that the excitation function contains a natural frequency, say
fit) = e
at
(9.40)
xf = Ae°' (9.41)
= e"
dx f
dt
-A(at + l)e°
d2 xf
2
= A(a 2 t + 2a)e a
dt
a — b
x = A.e"' + A 2 eb '
+ -^~
a — b
vg = 6e~ 2 '
+ 32
<L£
^F +
, i
10
r\
^ + 16* = 12er
Ci-X
~di
2'
+ 64 (9.44)
xn =A t
e- 2 ' + A 2 e-»'
Noting that the right member of the differential equation has the term e~ 2 ' in com-
mon with x„, we try
xf = Ate~ 2 '
+B
The factor t has been inserted into the natural trial solution of xf to remove the dupli-
cation of the term e~ 2 '. Substituting xf into (9.44) and simplifying, we have
6Ae~ 2 '
+ 165= 12e" 2 ' + 64
Chap. 9 Second-Order Circuits 217
x, = 2te" 2 ' + 4
As a final example, let us consider the case of (9.38) where b = a and/(f) is given
by (9.40). That is.
d2 x ~ dx
2a =£ + a 2x = e
a'
(9.45)
dt- dt
s1 — 2as + a2 =
and thus the natural frequencies are
st = s2 = a
x„ = (A, + A lt )e«
duplicated in the natural response. In this case, (9.43) will not work either because it,
too, is duplicated. The lowest power of / that is not duplicated is 2; thus we are led
to try
xf = At 2 e a '
EXERCISES
9.6.1 Find the forced response if
g+ 4 ^ + 3*=/(0
where /(/) is given by (a) 2e~ 3 ' — e~ 2 ' and (b) 4e~' + 2e~ 3 '.
W2+A Tt + <*=/(')
where f(t) is given by (a) 6e~ 2 ' and (b) 6te~ 2t . [Suggestion: In (b), try
xf = At 3 e~ 2 '.] Am. (a) 3t 2 e~ 2 ', (b) Oe~ 2t
9.6.3 Find the complete response if
d 2x
-ifi + 4x = 8 sin It
In the previous sections we have noted that the complete response of a circuit is the
sum of a natural and a forced response and that the natural response, and thus the
complete response, contains arbitrary constants. These constants, as in the first-order
cases of Chapter 8, are determined so that the complete response satisfies specified
initial energy-storage conditions.
To illustrate the procedure, let us find x(7), for t > 0, which satisfies the system
of equations
^ + 2x +
dt
5 ['
Jo
xdt= 16<r 3 '
(9.46)
x(0) = 2
To begin, let us differentiate the first of these equations to eliminate the integral; this
results in
^ + 2^ + 5x=
dt
1
dt
-48<r 3f
s
2
+ 2s +5=
with roots
x„ = e~\A cos x
2t +A 2 sin 2t)
xf = Ae~ 3 '
Chap. 9 Second-Order Circuits 219
we see that
x(t) = e-'(A {
cos It +A 2
sin 2t) - 6c 3 '
(9.47)
To determine the arbitrary constants we need two initial conditions. One, x(0) = 2,
is given in (9.46). To obtain the other we may evaluate the first of (9.46) at t = 0,
resulting in
•o
d
-^ + 2x(0) + 5 f x dt = 16
Noting the value of a(0) and that the integral term is zero, we have
dx(0)
-12 (9.48)
dt
x(0) =A -6= ! 2
or A :
— 8. To apply (9.48) we may differentiate (9.47), obtaining
-e~%A x
cos 2? +A 2 sin 2t) + 18e- 3 '
from which
^^ = 2A 2
-A + t
18 = 12 (9.49)
At this point let us digress for a moment to note a very easy way to get (9.49).
We may differentiate x(t) and immediately replace t by before we write down the
result. That is, in (9.47), the derivative of x at t = is e~' at t = (which is 1) times
the derivative of (A }
cos 2t +A 2
sin 2r) at f = (which is 2A 2 ) plus (^ cos 2/ +
^2 sin 2r)at t = (which is ^,) times the derivative at / (which is l)plus off = —
the derivative of — 6e~ 3 '
at t = (which is 18). These steps are written down in (9.49)
and can be done mentally, avoiding the intermediate prior step.
Returning to our problem, we now have the arbitrary constants, so that by (9.47),
the final answer is
As a last example, let us find v, t > 0, in the circuit of Fig. 9.5 if v^O) = v(0) =
and vg = 5 cos 2000? V. The nodal equation at node v {
is
220 Second- Order Circuits Chap. 9
o v
or
4u, 1
- v + 2 x 10- 3 $i = 2v gK = 10 cos 2000/ (9.50)
dt
10" 6 dv
\ x 10- 3 v, +\ X ^=
dt
or
v - _i x 10" 3 ^ (9.51)
^ + 2x
dt
2
10 3 ^
dt
+2x 10 6 v = -2 X 10 7 cos 2000/
i
2
+ 2x 10 3 s + 2 x 10 6 =
so that the natural frequencies are j 1>2 = 1000( — 1 ±yl). The natural response is
therefore
v. = e~ '{A x
cos 1000/ +A 2 sin 1000/)
(-2/4 + AB) cos 2000/ + (-4 A - 2B) sin 2000/ = -20 cos 2000/
Therefore equating coefficients of like terms, we have
Chap. 9 Second-Order Circuits 221
-2/4 + 45 = -20
-4/4 - 25 =
v = e-
1000
'^, cos 1000/ j A2 sin 1000/) + 2 cos 2000/ - 4 sin 2000/ (9.52)
= -{x
r,(o
+
) io- 3
^p
and since r,(0 + ) = ^
,
1
(0~) = we have
**P = (9.53)
t> = e- I000 '(-2 cos 1000/ + 6 sin 1000/) + 2 cos 2000/ - 4 sin 2000/ V
EXERCISES
9.7.1 Find x, t > 0, where
Jo
/o
x(0) = -1
u(t,(f)
EXERCISE 9.7.2
9.7.3 For t > in Ex. 9.7.2, find (a) v and (b) /'.
(Suggestion: Because of Kirchhoff's
laws and the terminal relationships of the elements, i has the same natural
frequencies as v. Thus v„ is easily obtained after i„ is found; its forced response
is evident by inspection of the circuit.)
Ans. (a) e _2 '(-6 cos At + 7 sin At) + 6, (b) e
-2 '(-2
cos 4/ — { sin 4r) + 2
One of the most important second-order circuits is the parallel RLC circuit, the source-
free version of which is shown in Fig. 9.6. We shall assume that at t = there is an
initial inductor current,
/(0) =h (9.55)
v(0) = V (9.56)
.
+ + /l + c *_o (9.57)
^.J.*
Chap. 9 Second-Order Circuits 223
Since there is no forcing function, the natural response is the complete response.
The characteristic equation is
As in the general second-order case already discussed, there are three types of
responses, depending on the nature of the discriminant, (\f2RC) 1 — (l/LC), in (9.59).
We shall now look briefly at these three cases.
Overdamped Case
L > 4R Z C (9.60)
then the natural frequencies of (9.59) are real and distinct negative numbers. In this
case, known as the overdamped ease, the response has the form
v = A.e 1 " + A 2e s *
(9.61)
dv(0 + ) _ V + RI
(9.62)
dt RC
which together with (9.56) can be used to determine the arbitrary constants.
As an example, suppose R = 1 Q, L — $ H, C — \ F, VQ = 2 V, and I = —3 A.
Then by (9.59) we have s, 2 = 1, —3, and hence —
v =A x
e~' + A 2 e~ 3'
v(0) = 2V
*£2 = 4v/s '
dt
v = 5e~' - 3e' 3!
This overdamped case is easily sketched, as shown by the solid line of Fig. 9.7, by
sketching the two components and adding them graphically.
The reason for the term overdamped may be seen from the absence of oscillations
(fluctuations in sign). The element values are such as to "damp out" any oscillatory
tendencies. Il is, of course, possible for the response to change signs once, depending
on the initial conditions.
L < 4R C 2
(9.63)
then the circuit is said to be underdamped. In this case the natural frequencies are
complex, and the response contains sines and cosines, which of course are oscillatory-
type functions. In this case it is convenient to define a resonant frequency,
COn (9.64)
JLC
a damping coefficient,
1
(9.65)
2RC
Second-Order Circuits 225
Each of these is a dimensionless quantity "per second." The resonant and damped
frequencies are defined to be radians per second (rad/s) and the damping coefficient
is nepers per second (Np/s).
Using these definitions, the natural frequencies, by (9.59), are
s lw2 = -a±ja>d
and therefore the response is
v = e~"(A 1
cos co d t + A2 sin co d t) (9.67)
From the initial conditions we have r(0) = and dv(0 + )/dt = 15 V/s, from which
A = x
and A 2 = 5. Therefore the underdamped response is
v = 5e~' sin 3/
This response is readily sketched if it is observed that since sin 3/ varies between
+1, and —l,v must be a sinusoid that varies between 5e~' and — 5<r'. The response
is shown in Fig. 9.8, where it may be seen that it is oscillatory in nature. The response
goes through zero at the points where the sinusoid is zero, which is determined, in
general, by the damped frequency co d .
L = 4R 2 C (9.68)
In this case the natural frequencies are real and equal, given by
5 lf2
= — cc, — oc
v = (A l
+ A z t)e-« (9.69)
v = Ate' 1 '
This is easily sketched by plotting At and e~ 2 and multiplying the two together. The
'
I 2
EXERCISES
9.8.1 In a source-free parallel RLC circuit, R = 1 kfi and C = 1 /^F. Find L so that
the circuit is overdamped with s 1>2 = —250, —750 s" (b) underdamped
(a) 1
,
9.8.2 (a) Find the differential equation satisfied by i in Fig. 9.6. (b) Use this result to
find i, for t > 0, if R= 5 CI, L= 1 H, C- 0.1 F, v(0) = 0, and ;(0) = 3 A.
d2i 1 di
Ans. (a) ^3 -f -g£ -jj +
1
jj^i = 0, (b) <r'(3 cos 3f + sin 30
9.8.3 The larger the value of R, the less damping there is in the underdamped case of
the parallel RLC circuit (because a = 1/2RC). Let R = oo (open circuit) and
show that
d 2v
dt 2
+ <olv =
The source-free series RLC circuit, shown in Fig. 9.10. is the dual of the parallel cir-
cuit, considered in the previous section. Therefore all the results for the parallel circuit
have dual counterparts for the series circuit, which may be written down by inspection.
In this section we shall simply list these results using duality and leave to the reader
their verification by conventional means.
cuit
v(0) = V
/(0) = /„
L §+ m+:vi"h+v'- (9.70)
Ls 2 + Rs + -1 = (9.71)
228 Second-Order Circuits Chap. 9
2
(9-72)
*-=-ffi±V(I) -ie
The series RLC circuit is overdamped if
AL
C> R2
(9.73)
i = A x
e "
s
+A 2e
s " (9.74)
C = AL2 (9.75)
R
in which case sx =s = 2
—R/2L, and the response is
i={A x
+A % t)e>* (9.76)
C<f 2
(9.77)
w = « -4=pr (9.78)
JLC
the damping coefficient is
a 4
=~2L ( 9 - 79 )
co d = jToi—^ (9.80)
Since the series and parallel RLC circuits are duals, we shall not consider any
examples in this section. In the following section we shall illustrate the series circuit
EXERCISES
9.9.1 Let R = 4Q, L - 1 H, r(0) = 4 V, and ;(0) = 2 A in Fig. 9.10. Find i for
/ > if C is (a) J^ f, (b) | F, and (c) ->
F.
Am. (a) 2e" 2 '(cos At - sin 4?) A, (b) 2(1 - 40c 2 '
A, (c) -3e"' + 5c A 3'
EXERCISE 9.9.2
In this section we shall consider the series RLC circuit excited by a source, an example
of which is shown in Fig. 9.1 1. The parallel RLC circuit is the dual circuit and we
shall leave examples of it to the problems.
2A
cuit
where the natural response v„ contains the natural frequencies. The natural frequencies
of the current i are the same as those of v because obtaining one from the other, in
230 Second-Order Circuits Chap. 9
general, requires only KirchhofTs laws and the operations of addition, subtraction,
multiplication by constants, integration, and differentiation. None of these operations
changes the natural frequencies. Therefore, since the natural frequencies of i are easier
to get (only one loop equation is required), let us obtain them. Around the loop we
have
j- + 2i + 5 ( idt + 6= 10 (9.82)
s
2
+ 2s + 5 =
with roots
*i,2 = —1 ±j2
Thus we have
v„ = e~'(A l
cos It + A2 sin It)
[We could have obtained s ul from (9.72) directly, but it is probably easier to write
the characteristic equation, since it can be done by inspection.]
The forced response vf is a constant in this case and may be obtained by inspection
of the circuit in the steady state. Since in the steady state the capacitor is an open
circuit and the inductor is a short circuit, if = and vf = 10. Therefore the complete
response is
v = e~'(A 1
cos It +A 2
sin It) + 10
v(0) = 6 =A + l
10
or A x
— —4. Also, we have
l*£2 = /(0) = 2
5 dt
or
2 1
dt
vg = 4 cos t V
In this case we shall need the differential equation for v, which we may get from (9.82)
and
Chap. 9 Second-Order Circuits 231
'-7$ (983>
We may also obtain it directly from the figure since the voltages across the inductor
and resistor are
&_ \_dH
1
dt 5 dt
and
2 dv
-,.
2i = -C--J-
5 dt
and that across the capacitor is v, of course. In either case, the result is
d2 v ~ dv c = 1rt
+
. .
diT+ dt
5v 20cost
The natural response is the same as in the previous example. To get the forced
response we shall try
vf = A cos t +B sin t
A = 4, B= 2
v = e~'(A 1
cos 2t + A2 sin It) + 4 cos t + 2 sin t
1,(0)= 6 = A x
+4
or A x
= 2. From the initial current and (9.83), we have
*W11= m = 2A 2z - A, '
+ 2
dt
or A2 = 5.
In the case of a driven, parallel RLC circuit, the complete response is obtained in
the same manner as that of the series case. Examples are given in Ex. 9.10.1 and in
the problems at the end of the chapter.
232 Second- Order Circuits Chap. 9
EXERCISES
9.10.1 Find / for t > if (a) L = f H and (b) L = 2 H.
Ans. (a) -3e-' + e' + 3'
2, (b) (-2 - 4t)e~ 2 ' + 2
EXERCISE 9.10.1
EXERCISE 9.10.2
We shall close this chapter by considering two methods of expediting the process of
obtaining the describing equation of the circuit. In the case of the parallel and series
RLC circuits, a single equation is required, which, after differentiation, is the describ-
ing equation. However, in many second-order circuits there are two simultaneous
circuit equations from which the describing equation is obtained after a tedious
elimination process.
As an example, let us consider the circuit of Fig. 9.12 for / > 0. Taking node b
as reference and writing nodal equations at nodes a and v we have x
v- v*
|
v — Vi i
"*"
_L* = o
"t"
4 '
" 6 "
4 dt
(9.84)
v
-^=^ + v, dt + /(0) -
J"
Chap. 9 Second-Order Circuits 233
If we are interested in finding v we must eliminate r, and obtain the describing equation
in terms of v. The result, as the reader may verify, is
dt
1
dt
+ dt ^ °
(9.85)
In this case the process is not overly complicated but it can be shortened by the
dt
It is important here to note that x is factored out of the middle member and placed
after the operator expression, indicating that the operation is to be performed on x.
/ 1
n ,
5\ 1 1
-4i»+0+i)*,=o
which when cleared of fractions becomes
+ 5)v — 2v = 3v g
(3D l
(9.86)
-Dv + (D + 6)v, =0
To eliminate v u let us "operate" on the first equation with D+ 6, by which we mean
234 Second-Order Circuits Chap. 9
multiply it through by D+ 6 on the left of each term. Then let us multiply the second
equation by 2, resulting in
Multiplying the operators as if they were polynomials, collecting terms, and dividing
out the common factor 3, we have
The procedure may be carried out in a more direct manner by using determinants.
For example, we may use Cramer's rule to obtain the expression for v from (9.86),
given by
(9.88)
3D + 5 -2
A
1\ — (9.89)
-D D+6
and Aj is given by
3v g -2
A,= = 3(ZM (9.90)
D+6
We note that in this last expression we must be careful to write v g after the operator.
Writing (9.88) as
Av = Aj
we see from (9.89) and (9.90) that we have the describing equation (9.87).
The second method we shall consider is a mixture of the loop and'nodal methods
in which we select the inductor currents and the capacitor voltages as the unknowns,
rather than the loop currents or the node voltages. We then write KVL around loops
which contain only a and KCL at nodes, or generalized nodes, to which
single inductor
only a single capacitor is connected. In this manner each equation contains only one
derivative, that of an inductor current or a capacitor voltage, and no integrals. The
equations are then relatively easy to manipulate to find the describing equation.
As an example using Fig. 9. 12 again, let i, the inductor current, and v, the capacitor
voltage, be the unknowns. (These unknowns are sometimes called the state variables
of the circuit.) Then the nodal equation at node a is
Chap. 9 Second-order Circuits 235
v = 6/ +
di_
(9.92)
dt
It is a relatively simple matter to solve for i in (9.91), substitute its value into (9.92),
and simplify the result to obtain (9.85).
Some advantages of this method are that no integrals appear (thus no second
derivatives occur as a result of differentiation), one unknown is easily found in terms
of the others, and the initial conditions on the first derivatives are easily obtained for
use in determining the arbitrary constants in the general solution. For example, it
dv(0 + )
v g (0 + ) - 10
dt
This last method can be greatly facilitated, particularly in the case of complex
circuits,by using graph theory. Since we are looking for inductor currents and capac-
itor voltages (the state variables), we put the inductors in the links, whose currents
constitute an independent set, and the capacitors in the tree, whose branch voltages
constitute an independent set, as we recall from Chapter 6. (The tree should also
contain voltage sources, and the links should contain current sources, etc., if possible.)
Each inductor L is then a link with current which forms a loop whose only other/',
elements are tree branches. Therefore KVL around this loop will contain only one
derivative term, L(di/dt), and no integrals. This loop can easily be found since it is
the only loop in the graph if the only link added to the tree is L. For example, the
graph of Fig. 9.12 is shown in Fig. 9.13 with tree branches shown as solid lines and
links as dashed lines. The loop containing the 1-H inductor is a,v u b,a through the
branch labeled v, and KVL around it is (9.92).
Each capacitor is a tree branch whose current, together with link currents, consti-
tutes a set of currents flowing out of a node or a generalized node, because if the capac-
from the circuit, the tree is separated into two parts connected only by links.
itor is cut
In the example of Fig. 9.13 the three currents, shown crossing the dotted line through
the capacitor, labeled v, and two links, add to zero by KCL. This is stated by (9.91).
EXERCISES
9.11.1 Find the describing equation for v, t > 0, in Ex. 9.10.2 by (a) the first method
of this section using nodal equations and (b) the second method of this section.
Ans. CidHldf1 ) + 6C(dvldt) + v = 24
9.11.2 Solve Ex. 9.1.1 using (a) the first method of this section applied to the mesh
equations and (b) the second method of this section.
PROBLEMS
9.1 Find z'i and i 2 , for t > 0, in Ex. 9.1.1 if vg = (a short circuit), j'i(0)
= 2 A,
and / 2 (0) = 3 A.
PROBLEM 9.3
9.4 The circuit is in steady state when the switch is opened at / = 0. Find v and i
for t > 0.
9.5 Insert a 1-fi resistor in series with v g in Fig. 9.2, thereby making the source a
practical rather than an ideal one. Show that in this case v 2 satisfies the second-
order equation
5^ + ii^l +4 =4^ +
dt< dt
,2
dt
4f ;
Chap. 9 Second-Order Circuits 237
t = o
16V
PROBLEM 9.4
PROBLEM 9.6
9.7 Find v, for r > 0, if the circuit is in steady state when the switch is moved from
position 1 to position 2 at / = and L is (a) 2 H, (b) 4 H, and (c) J^2 H.
— 8V
PROBLEM 9.7
9.8 Find ;', for t > 0, if the circuit is in steady state when the switch is opened at
t = 0.
PROBLEM 9.8
238 Second-Order Circuits Chap. 9
PROBLEM 9.10
9.11 Find v, for t > 0, if the circuit is in steady state when the switch is moved from
position 1 to position 2 at t = 0.
I kn
—2 o \ o—
I
h
v "
t =
10 V O.I H 5 V
0.025 gF t
PROBLEM 9.11
PROBLEM 9.12
9.13 Find /, / > 0, if the circuit is in steady state when the switch is opened at t = 0.
t =o.
PROBLEM 9.13
Second-Order Circuits 239
2u(t) A
€^
2A IH
l
g>
e-
6u(t) V
—
PROBLEM 9.14
9.15 Find v, t > 0, in the figure for Prob. 9.7 if L — 1 H and the switch is moved from
position 2 to position 1 at / = 0.
_t
3e A
2H J_ 2H
M\ —TSV> *"
i
2A F;±: v
PROBLEM 9.16
9.17 Find the maximum value of the overdamped response of Prob. 9.7(c) and the
time at which it occurs.
9.18 Find /, t > 0, if there is no initial stored energy and (a) R = \ Q, // = 2; (b)
r = i.
Q, ix = 1 ; and (c) R = \ Q, fj.
= 2.
2Q
PROBLEM 9.18
9.19 Find /', / > 0, if there is no initial stored energy and (a) C= \ F, (b) C= \ F,
and (c) C= TV$ F.
240 Second-Order Circuits Chap. 9
v,/2 A
PROBLEM 9.19
9.20 Find v, t > 0, in the circuit of Fig. 9.12 if vg = 29 cos 2t V and the circuit is in
steady state when the switch is moved at t = 0.
2d 3fl
PROBLEM 9.21
.I o v
PROBLEM 9.22
9.24 Find 0, if (a) ^(0) = 4 V, «(0) = 0; (b) y,(0) = 0, i>(0) = 2 V; and (c)
v, t >
i-,(0) 4 V, v(0) = 2 V. (Note that the response is an unforced sinusoidal
=
response. Such a circuit is called a harmonic oscillator.)
Chap. 9 Second-Order Circuits 241
in
JSA^r-
in
in
I F 1.
I '
3
n
in
PROBLEM 9.23
in
-\KAr-
m
in
1 I F
m
_+
PROBLEM 9.24
9.25 Higher-order differential equations may be solved in the same manner as second-
order equations. There are more natural frequencies and thus more terms in the
natural response. For example, if
d3x , d 2x dx .-
dF +
.
6
-d^
.,.
+ n ^t
+6x =
. ,
l2
s3 + 6s 1 + \\s + 6 =
s= -1, -2, -3
242 Second-Order Circuits Chap. 9
xn =A x
e~' + A 2 e- 2 + '
A^<
Show also that the forced response is
Xf = 2
x = xn + xf
9.26 Using the results of Prob. 9.25, find i, t > 0, if there is no initial stored energy.
^o
PROBLEM 9.26
10
SINUSOIDAL EXCITATION
AND PHASORS
In the two previous chapters we have analyzed circuits containing dynamic elements
and have seen that the complete response is the sum of a natural and a forced response.
The natural response for a given circuit is obtained from the dead circuit and therefore
is independent of the sources, or excitations. The forced response, on the other hand,
depends directly on the type of excitation applied to the circuit. In the case of a dc
source, the forced response is a dc steady-state response, an exponential input results
in an exponential forced response, and so forth.
One of the most important excitations is the sinusoidal forcing function. Sinusoids
abound in nature, as, for example, in the motion of a pendulum, in the bouncing of a
ball, and in the vibrations of strings and membranes. Also, as we have seen, the
the ac steady-state response that is left after the short time required for the transitory
natural response to die.
Since we are interested only in the ac steady-state response, we shall not limit
ourselves, as we did in Chapters 8 and 9, to first- and second-order circuits. As we shall
243
244 Sinusoidal Excitation and Phasors Chap. 10
see, higher-order RLC circuits may be handled, insofar as the ac steady-state response
is concerned, in the same way as resistive circuits.
which is sketched in Fig. lO.l. The amplitude of the sinusoid is Vm which is the ,
maximum value that the function attains. The radian frequency, or angular frequency,
is co, measured in radians per second (rad/s).
The sinusoid is aperiodic function, defined generally by the property
where T is the period. That is, the function goes through a complete cycle, or period,
which is then repeated, every T seconds. In the case of the sinusoid, the period is
2n
(10.3)
CO
as may be seen from (10. 1) and (10.2). Thus in 1 s the function goes through l/T
cycles, or periods. Its frequency f is then
co
/ =
T In
(10.4)
i
v(t)'
olt7
cycle
cycles per second, or hertz (abbreviated Hz). The latter term, named for the German
physicist Heinrich R. Hertz ( 1 857-1894), is now the standard unit for frequency. Some
older books use the former term, but it is being discontinued. The relation between
frequency and radian frequency is seen by (10.4) to be
to = 2nf (10.5)
where is the phase angle, or simply the phase. To be consistent, since cot is in radians,
should be expressed in radians. However, in electrical engineering it is often con-
\enient to specify in degrees. For example, we may write
v = Vm sin (* +
*)
or
As an example, consider
v, = 4sin(2r + 30°)
and
v2 = 6sin(2r - 12°)
Thus far we have considered sine functions rather than cosine functions in defining
A
or
in (cot
sin + y \ = cos cot (10.8)
The only difference between sines and cosines is thus the phase angle. For example,
we may write (10.6) as
To determine how much one sinusoid leads or lags another of the same frequency,
we must first express both as sine waves or as cosine waves with positive amplitudes.
For example, let
v x
=4 cos (It + 30°)
and
Then, since
we have
vz = 2sin(2r + 18° + 180°)
= 2 cos (It + 18° + 180° - 90°)
= 2cos(2r + 108°)
Comparing this last expression with v ,, we see that v x leads v 2 by 30° — 108° = —78°,
which is the same as saying that i\ lags v z by 78°.
The sum of a sine wave and a cosine wave of the same frequency is another
sinusoid of that frequency. To show this, consider
B
A cos cot + B sin cot = J z
+ B'
V^ + B
2 2 ^
cos cat ^
'
JA ,
,,1
+ B„,
,
2
sin cot
vlt ) 1
Vm sin (cott<J>)
//% //
Vm sin u) t
s&//
9>
A x
s\ 9
FIGURE 10.2 Two sinusoids with different phases FIGURE 10.3 Triangle useful
in adding two sinusoids
= tan'4 (10.10)
A
A similar result may be established if the sine and cosine terms have phase angles
other than zero, indicating that, in general, the sum of two sinusoids of a given
frequency is another sinusoid of the same frequency.
As an example, we have
12
-5 cos + = J5 + -
3t 12 sin 3r 1
12 2 cos 3/ tan" 1
m.
= 13 cos (3/ - 112.6°)
since tan
-1
(12/— 5) is in the second quadrant, because A = —5 < and B — 12 > 0.
EXERCISES jCl, /
c
10.1.1 Find the period of the following sinusoids:
(a) 6sin(5r + 17°) f.
SOI
(b) cos (2/ + + 2 sin (it - |-).
-J)
(c) 4 cos In t. Ans. (a) 271/5, (b) n, (c) 1
As an example of a circuit with a sinusoidal excitation, let us find the forced com-
ponent if of the current /'
in Fig. 10.4. The describing equation is
and, following the method of the previous chapter, let us assume the trial solution
L{— coA sin cot + coB cos cot) + R(A cos cot + B sin cot) = Vm cos cot
Therefore equating coefficients of like terms, we must have
RA + = Vm
coLB
-coLA + RB =
from which
= RV„
A
R2 + co
2
L2
coLVm
B=
R2 + co
2
L2
The forced response is then
if
f
=
R2 +
RV„
co
2
L 2 ^
cos ""
cot — -^-
i? + 2
—L co
2__
2 2
Si n at
V. ,coL
'/ cos cot — tan (10.12)
JR + co L 2 2
R)
Vm cos cot
The forced response is, therefore, a sinusoid like the excitation, as we predicted when
we chose the trial solution. We may write it in the form
where
L=
V-R 2 + co
2
L 2
and
<t>
= -tan-^ (10.14)
i„ = A l
e'
R,L
EXERCISES
10.2.1 Find the forced response i
f in Fig. 10.4 if L = 4 mH, R = 3 kfi, Vm = 5 V,
and co = 10 6 rad/s. Ans. cos (10 6 r - 53.1°) mA
10.2.2 Find the forced component of v.
EXERCISE 10.2.2
250 Sinusoidal Excitation and Phasors Chap. 10
numbers. The reader who is unfamiliar with complex numbers, or who needs to review
the subject, should consult Appendices C and D, where complex numbers and their
properties are discussed in some detail. For convenience, we shall list a few of these
properties in this section before considering the alternative method of analysis.
The complex number A is written in rectangular form as
A = a+jb (10.15)
where y = <J — 1 and the real numbers a and b are the real and imaginary parts of A,
respectively. Equivalently we may say that
a = Re A, b = Im A
where Re and Im denote "the real part of" and "the imaginary part of."
The number A may be written also in the polar form,
where \A |
is the magnitude, given by
a = tan"
a
These relations between rectangular and polar forms are illustrated in Fig. 10.5.
As an example, suppose we have A — A +7*3. Then \A |
= *J4
2
+ 3
2
= 5 and
a = tan
-1
| = 36.9°. Therefore the polar form is
A = 5 /36.9°
As another example, consider A = —5 — jl2. Since both a and b are negative, the
line segment representing A lies in the third quadrant, as shown in Fig. 10.6, from
which we see that
\A\ = V5 + 2
12 2 = 13
and
a= 180° + tan-'-^ = 247.4°
A= o + |b
FIGURE 10.5 Geometrical representation of FIGURE 10.6 Complex number with negative
a complex number A real and imaginary parts
j= /90° 1
f= - = 1 1/180"
and so forth.
By Euler*s formula, which is discussed in Appendix D and which we have used
previously in Chapter 9, we know that
and
Returning to the RL circuit example of Fig. 10.4, we know that exponentials are
mathematically easier to handle as excitations than sinusoids. Therefore let us see
what happens if we apply the complex excitation
y e
j<»<
(10.18)
ve = Vm cos cot = Re v x
(10.19)
i
f
= Ae' u
from which
A =
R + jcoL
-
j tan ' coL/R
JR. 1 + co 2
L2
;'(cu(-tan-' 01L! R)
JR + co 2 2
L 2
WR 2
+ o>
2
L2
Vm —
JR. 2
+ co
-^-^ cos (cot
2
L2 \
tan 1
%R)
r )
which, by (10.12), is the correct forced response of Fig. 10.4. That is,
i
f
= Re /j (10.21)
We have established for this example the interesting result that if/, is the complex
response to the complex forcing function v u then i
f
= Re/, the response to is vg =
Re v1 . That is, v x
yields /, and Re v x
= v g yields Re /, = f The reason for this
i . is that
the describing equation (10.20) contains only real coefficients. Thus, from (10.20), we
have
di
Re^Ly + Ri\ = Rev,
or
i = i
f
= Re/, (10.22)
Thus we see that it is easier to use the complex forcing function i\ to find the
complex response /,. Then since the real forcing function is Re v u the real response is
Re /,. This principle holds for all our circuit analyses, since the describing equations
are linear with real coefficients, as may be seen in a development analogous to that
leading to (10.22).
EXERCISES
10.3.1 Replace the real forcing function Im cos cot in Ex. 10.2.2 by the complex forcing
function Im e Ja", find the resulting complex response v u and show that the real
response is v = Re vx .
H % =a &
a
dx\ d
{K ' x)
and use this result to establish (10.22). {Suggestion: Let x = f + jg, where/
and g are real.)
Let us now
generalize the results using complex excitation functions in the previous
section.The excitation, as well as the forced response, may be a sinusoidal voltage or
current. However, to be specific, let us consider the input to be a voltage source and
the output to be a current through some element. The other cases may be considered
in an analogous way.
In general, we know that if
Vg = Vm cos(tot + 6) (10.23)
as indicated in the general circuit of Fig. 10.7. Therefore if by some means we can find
Im and 0, we have our answer, since co, 9, and Vm are known.
,
Vq = Vm COSlcot +9) j
i
= Im cos(cot +<P)
v, = Vm e J{a" +e)
(10.25)
and find the complex response i u as shown in Fig. 10.8. Then we know from the
results of the previous section that the real response of Fig. 10.7 is
i = Re / (10.26)
This is a consequence of the fact that the coefficients in the describing equation are
real, as pointed out previously.
j (co1 1 )
v, = Vm e
The describing equation may be solved for the forced response by the method of
Chapter 9. That is, since we may write the excitation as
v x
= Vm e je
e i0
"
(10.27)
/, = Ae icot
A = Im e j *
Chap. 10 Sinusoidal Excitation and Phasors 255
and hence
i, = Im e J +e JO
"
(10.28)
d2 i =
dT2 , +
,
^
r.
2
di_
^
dl
+ 8/ 12^/T cos (2f + 15°)
+15t
^1 + 2^+8/,= 12 x/T^' (2
' >
or
A
n^/T^ _ 12VT/15_°
~
4+y4 4^T/45° ~ ^^-^
which gives
/, = (3 /-30V 2'
_ 3 e ;(2t-30°)
Thus the real answer is
i
f
= Re/, = 3 cos (2?- 30°)
EXERCISES
10.4.1 Find the forced response v using the method of this section.
17 cosl6t V v
EXERCISE 10.4.1
10.4.2 Find the forced response i using the method of this section if vg = 6 sin At.
[Suggestion : sin At = Im e'*'] Ans. 3 sin At A
I H 2 n
->m-
© ZZ 1/16 F
EXERCISE 10.4.2
10.5 PHASORS
The results obtained in the previous section may be put in much more compact form
by the use of quantities called phasors, which we shall introduce in this section. The
phasor method of analyzing circuits is credited generally to Charles Proteus Steinmetz
(1865-1923), a famous electrical engineer with the General Electric Company in the
early part of this century.
To begin, let us recall the general sinusoidal voltage,
which, of course, is the source voltage v g of the previous section. If the frequency co is
known, then v is completely specified by its amplitude Vm and its phase 9. These
quantities are displayed in a related complex number,
V = Vn e» = VJ9 (10.30)
w = Re(Ve**) (10.32)
V= 10/30_° V
since Vm =
10 and 9 = 30°. Conversely, since co = 4 rad/s is assumed to be known,
v is readily obtained from V.
In an identical fashion we define the phasor representation of the time-domain
current
to be
\ = Im e* = lJ± (10.34)
Thus if we know, for example, that co = 6 rad/s and that I = 2/15° A, then we have
V= 8 /-60
To see how the use of phasors can greatly shorten the work, let us reconsider Fig.
10.4 and its describing equation (10.11), rewritten as
Following our method, we replace the excitation Vm cos cot by the complex forcing
function
Vl = Vm e^'
,
Vl = \e Ja"
i = Re /
Next, trying
/j = Ie
/»>'
as a solution, we have
jcoLle im! + ja
/?Ie " = Ve- Q"
'
jcoLl + Rl =V (10.36)
Therefore
K- 1
'-tan-*R
/? + ;o>L Jri + co
2
L2 1
it = ,Jloit-tan-' coLIR)
JR 1
+(o 2 L 2
vast saving of time and effort. Also, in the process we have converted the differential
equation into an algebraic equation, somewhat like those encountered in resistive
circuits. Indeed, the only difference is that the numbers here are complex, whereas in
resistive circuits they were real. With the hand calculator as commonly available as it
is today, even the complexity of the numbers presents little difficulty.
In the remainder of the chapter we shall see how to bypass all the steps between
(10.35) and (10.36) by studying the phasor relationships of the circuit elements and
considering Kirchhoff's laws as they pertain to phasors. Indeed, as we shall see, we
may go directly from the circuit to (10.36), bypassing even the step of writing down the
differential equation.
In general, the real solutions are time-domain functions, and their phasors are
frequency-domain functions; i.e., they are functions of the frequency a. This is illus-
trated by the phasor I of the last example. Thus to solve the time-domain problems, we
Chap. 10 Sinusoidal Excitation and Phasors 259
EXERCISES
10.5.1 Find the phasor representation of (a) 10 cos {It -f 45°), (b) 4 cos 2/ + 3 sin
It, and (c) -6 sin (5f - 75°).
10.5.2 Find the time-domain function represented by the phasors (a) 10 / — 12° , (b)
6 +j8, and (c) —jA. In all cases co = 3.
Ans. (a) 10 cos (3* - 12°), (b) 10 cos (3? + 53.1°), (c) 4 cos (3f - 90°)
In this section we show that relations between phasor voltage and phasor current
shall
for resistors, inductors, and capacitors are very similar to Ohm's law for resistors. In
fact, the phasor voltage is proportional to the phasor current, as in Ohm's law, with
the proportionality factor being a constant or a function of the frequency co.
v = Ri (10.37)
where
v = Vm cos (cot + 6)
(10.38)
i = Im cos (cot + 0)
+e)
If we apply the complex voltage Ym e Jico '
, the complex current which results is
ibat+
Im e *\ which substituted into (10.37) yields
+ B) __ +
y e
J(cot
ft J e
J(cot 4>)
Vm e je = RIm e» (10.39)
which, since Vm e' e and lm e'* are the phasors V and I, respectively, reduces to
V = R\ (10.40)
260 Sinusoidal Excitation and Phasors Chap. 10
Thus the phasor or frequency-domain relation for the resistor is exactly like the time-
domain relation. The voltage-current relations for the resistor are illustrated in
Fig. 10.9.
v = Ri V = RI
(a) (b)
is applied across a 5-Q resistor, with the polarity indicated in Fig. 10.9(a). Then the
phasor voltage is
v,i
This is. of course, simply the result we would have obtained using Ohm's law.
In the case of the inductor, substituting the complex current and voltage into the
time-domain relation,
di
v =L ,
dt
Again, dividing out the factor e ico and identifying the phasors, we obtain the phasor
'
relation
\=jcoLl (10.43)
Thus the phasor voltage V, as in Ohm's law, is proportional to the phasor current I,
with the proportionality factor jcoL. The voltage-current relations for the inductor are
shown in Fig. 10.1 1.
i I
O 1 o
+ +
(a) (b)
If the current in the inductor is given by the second of (10.38), then by (10.43), the
phasor voltage is
v = (M)(/J)
= coLIJtj) + 90°
Comparing this result with the second of (10.38), we see that in the case of an inductor
262 Sinusoidal Excitation and Phasors Chap. 10
the current lags the voltage by 90°. Another expression that is used is that the current
and voltage are 90° out of phase. This is shown graphically in Fig. 10.12.
V,l
Finally, let us consider the capacitor. Substituting the complex current and voltage
into the time-domain relation.
i = C dv
di
J
dt
= jcoCVm e Ho" +m
Again dividing by e
j0J '
and identifying the phasors, we obtain the phasor relation
I=jo>CV (10.44)
or
(10.45)
jcoC
Thus the phasor voltage V is proportional to the phasor current I, with the propor-
tionality factor given by 1/jcoC. The voltage-current relations for a capacitor in the
time and frequency domains are shown in Fig. 10.13.
In the general case, if the capacitor voltage is given by the first equation of (10.38),
then by (10.44), the phasor current is
/ = (jcoC)(VJ9)
= coCVje + 90°
Chap. 10 Sinusoidal Excitation and Phasors 263
i
= Cdv/dt 1= jujCV
jooC
(a) (b)
which, by comparison with the first equation of (10.38), indicates that in the case of a
capacitor the current and voltage are out of phase with the current leading the voltage
by 90°. This is shown graphically in Fig. 10.14.
v,i
I=y(100)(10-«)(10/30°)A
= 1 /120° mA
The time-domain current is then
EXERCISES
10.6.1 Using phasors, find the ac steady-state current iif v = 12 cos (1000/ + 30°) V
in (a) Fig. 10.9(a) for R = 4kft, (b) Fig. 10.11(a) for L = 20 mH, and (c)
Let us now consider a general phasor circuit with two accessible terminals, as shown in
Fig. 10.15. If the time-domain voltage and current at the terminals are given by
(10.38), then the phasor quantities at the terminals are
v = vje
— (10.46)
I = UA
We define the ratio of the phasor voltage to the phasor current as the impedance of
the circuit, which we denote by Z. That is,
(10.47)
i
which by (10.46) is
where |
Z |
is the magnitude and 6 Z the angle of Z. Evidently,
Impedance, as is seen from (10.47), plays the role, in a general circuit, of resistance in
resistive circuits. Indeed, (10.47) looks very much like Ohm's law; also like resistance,
impedance is ohms, being a ratio of volts to amperes.
measured in
It is important to stress that impedance is a complex number, being the ratio of two
complex numbers, but it is not a phasor. That is, it has no corresponding sinusoidal
time-domain function of any physical meaning, as current and voltage phasors have.
The impedance Z is written in polar form in (10.48); in rectangular form it is
Chap. 10 Sinusoidal Excitation and Phasors 265
generally denoted by
Z = R + jX (I0.49)
|Z| JR + X Z 1
ez tan -i
*
A'
and
7? = Z | |
cos Z
X= |Z|sin6> z
These relations are shown graphically in Fig. 10.16.
As an example, suppose in Fig. 10.15 that V = 10 /56.1° V and I = 2 /20° A. Then
we have
10/56.1°
= 5 / 36
jo
n
2/20°
ZR = R
Z L = joiL = coL/90°
(10.50)
PHASOR
V
CIRCUIT
In the case of a resistor, the impedance is purely resistive, its reactance being zero.
Impedances of inductors and capacitors are purely reactive, having zero resistive
components. The inductive reactance is denoted by
XL = coL (10.51)
so that
ZL =jXL
and thus
Z c =jXc (10.53)
Since co, L, and C are positive, we see that inductive reactance is positive and that
capacitive reactance is negative. In the general case of (10.49), we may have X— 0, in
which case the circuit appears to be resistive; X> 0, in which case its reactance
appears to be inductive; and X < 0, in which case its reactance appears to be capaci-
tive. These cases are possible when resistance, inductance, and capacitance are all
present in the circuit, as we shall see. As an example, the circuit with impedance given
by Z— 4 + y'3, which we have just considered, has reactance X = 3, which is of the
inductive type. In all cases of passive circuits, as we shall see in Chapter 12, the
resistance R is nonnegative.
The reciprocal of impedance, denoted
Y = -L (10.54)
Y = G+jB (10.55)
To obtain the relation between components of Y and Z we may rationalize the last
member of (10.56). which results in
R-jX
Rz + X 2
Equating real and imaginary parts results in
R
G
R1 + X 2
(10.57)
D _ X
R + X1 2
Therefore we note that R and G are not reciprocals except in the purely resistive case
{X — 0). Similarly, X and B are never reciprocals, but in the purely reactive case
(R = 0) they are negative reciprocals.
As an example, if we have
Z = 4+y3
then
v- ^__ 4-y3 _
42 + 2
4
25
J
J
3_
1?
25
;3 3
Therefore G— -^ and B =- —
__!_
^
25-
Further examples are
YR = G
YL --L
~jcoL
Y c = jcoC
which are the admittances of a resistor, with R— \]G, an inductor, and a capacitor.
EXERCISES
10.7.1 Find the impedance seen at the terminals of the source in Fig. 10.4 in both
rectangular and polar form.
Ans. R +j(OL, ^R + co
2 2
L 2 ltan~ coL/R
l
268 Sinusoidal Excitation and Phasors Chap. 10
10.7.2 Find the admittance seen at the terminals of the source in Fig. 10.4 in both
rectangular and polar form.
. R coL 1
/ _ coL
Ans Ri J
l
-
+ Oi'-U- Ri + C0 2 L 2 '
JR* + CQiL ! 2 R
10.7.3 Find the conductance and susceptance if Z is (a) 5 +7 12, (b) 0.8 — y'0.6, and
(c) ^2/45°. Ans (a)^,
hb, - J&,
--&, (b) 0.8, 0.6, (c) i,
{,
- -\
Kirchhoffs laws hold for phasors as well as for their corresponding time-domain
We may see this by observing that if a complex excitation, say
voltages or currents.
Vm e il0J,+e \ is applied to a circuit, then complex voltages, such as V 1
e Mo" +Bl) ,
V2 e Jlo" +e,-\ etc., appear across the elements in the circuit. Since Kirchhoffs laws hold
in the time domain, KVL applied around a typical loop results in an equation such as
y e
j(ait+e,) _j_ y e
j(mt+e 2 ) _i
, -\- V e
iio" +BN) =
Dividing out the common factor e io", we have
V, +v + 2 . . . + v„ =
where
VJd„ n = l,2,...,N
are the phasor voltages around the loop. Thus KVL holds for phasors. A similar
development will also establish KCL.
In circuits having sinusoidal excitations with a common frequency co, if we are
interested only in the forced, or ac steady-state, response, we may find the phasor
voltages or currents of every element and use Kirchhoffs laws to complete the
analysis. The ac steady-state analysis is therefore identical to the resistive circuit
analysis of Chapters 2, 4, and 5, with impedances replacing resistances and phasors
replacing time-domain quantities. Once we have found the phasors, we can convert
immediately to the time-domain sinusoidal answers.
As an example, consider the circuit of Fig. 10.17, which consists of N impedances
connected in series. By KCL for phasors, the single phasor current I flows in each
element. Therefore the voltages shown across each element are
V,=Z,I
v =Z I
2 2
Vjv — LN i
Chap. 10 Sinusoidal Excitation and Phasors 269
+ o-
*l
L
2 ^
+ + +
V — v ,
V2 VN
V = V, + V 2 + . . . + VN
= (Z, +Z + 2 ... +Z JV )I
V = Z eq I
Z. eq = Zq +Z 2 -p . . . -p ZN (10.58)
Y2 + . . . + Y„ (10.59)
1 1 Z]Z;
(10.60)
Y, +Y 2 Z, +Z 2
In like manner, voltage and current division rules hold for phasor circuits, with
impedances and frequency-domain quantities, in exactly the same way that they held
for resistive circuits, with resistances and time-domain quantities. The reader is asked
to establish these rules in Ex. 10.8.2.
For example, let us to return to the RL circuit considered in Sec. 10.2. The circuit
and its phasor counterpart are shown in Figs. 10.18(a) and (b), respectively. By KVL
ZJ + RI= VJ0
270 Sinusoidal Excitation and Phasors Chap. 10
(a) (b)
or
VJO
R + jcoL
iCoL
-tan
V-R 2
+ co
2
L2 ~R
cos ( -tan-^)
JR + a2 2
LI
"
Z R + jcoL
as obtained earlier.
EXERCISES
10.8.1 Derive (10.59).
Z 1
+Z 2
Chap. 10 Sinusoidal Excitation and Phasors 271
_ Y2 _ Z
Y,+Y I.=
¥ T 1
*« Z,+Z 2 2
+ o- o t
o •
(o) (b)
EXERCISE 10.8.2
3 n I H
AW 'TRRP-
+
EXERCISE 10.8.3
10.8.4 Find the steady-state value of v in Ex. 10.8.3 using voltage division. Check by
using the fact that the current leads the voltage by 90° in a capacitor.
Ans. 2 cos (4/ - 135°) V
As the discussion in the previous section suggests, we may omit the steps of finding the
describing equation in the time domain, replacing the excitations and responses by
their complex forcing functions and then dividing the equation through by eJa" to
obtain the phasor equation. We may simply start with the phasor circuit, which is the
time-domain circuit with the voltages and currents replaced by their phasors and the
elements identified by their impedances, as illustrated previously in Fig. 10.18(b). The
describing equation obtained from this circuit is then the phasor equation. Solving
this equation yields the phasor of the answer, which then may be converted to the time-
domain answer.
272 Sinusoidal Excitation and Phasors Chap. 10
The procedure from starting with the phasor circuit to obtaining the phasor
answer is identical to that used earlier in resistive circuits. The only difference is that
the numbers are complex.
As an example, let us find the steady-state current i in Fig. 10.19(a). The phasor
circuit, shown in Fig. 10.19(b), is obtained by replacing the voltage source and the
currents by their phasors and labeling the elements with their impedances. In the
phasor circuit the impedance seen from the source terminals is
7 _ i .
(3+73)(-j3)
= 4-j3Q
Therefore we have
5/0 5/0
I, 1/36.9°
4 - ;3 5/-36.9
c
/ = v^cos(3/ + 81.9°) A
1/9 F
(a)
5ZQ V ^ -j3 n
(b)
14 3/0 (10.61)
-J2
By Ohm's law we have V, = 41. which substituted into (10.61) yields
-j6 6 /-90
I
- )2 /-45°A
2 2y r 2~ /-45
Therefore we have
3
cos (4r - 45°) A
1/8 F
3 cos 4 t 1/2 v
(a)
]2 n
3/0 A 1/2 V,
(b)
In the case of an op amp. the phasor circuit is the same as the time-domain circuit.
That is. an ideal op amp in the time-domain circuit appears as an ideal op amp in the
phasor circuit, because the time-domain equations
i = 0, v =
which characterize the current into and the voltage across the input terminals retain
the identical form,
1 = 0, V=
in the phasor equations.
EXERCISES
10.9.1 Solve Ex. 10.2.2 by means of the phasor circuit.
10.9.2 Find the forced response v in the figure for Ex. 10.4.1 by means of the phasor
circuit.
10.9.3 Find the forced response i in the figure for Ex. 10.4.2 by means of the phasor
circuit.
PROBLEMS
10.1 Given
vi = 10 cos At
PROBLEM 10.2
Chap. 10 Sinusoidal Excitation and Phasors 275
10.3 In the figure of Ex. 10.2.2, if the source is 2 cos 5000/ mA and the output is
A. Find the current response to a voltage source of (a) \6e J{i "* s) V, (b)
8 cos (3/ + 35 ) V, and (c) 2 sin (3/ - 60 ) V.
10.7 Find the impedance of the circuit of Fig. 10.15 if the time-domain functions
represented by the phasors V and I are
(a) -5 cos It + 12 sin It V, = 1.3 cos (2/ + 40°) A.
v = i
0.05 H
4 cos 20,000 1 I kO
mA
PROBLEM 10.11
10.12 Find the power being delivered by the source in Prob. 10.11 at / = 7T/20 ms.
10.13 Show that if the capacitance in the figure for Prob. 10.1 1 is changed to 0.025 /zF,
then the impedance seen at the terminals of the source is real. In this case, find
the power being delivered by the source at / = fl/20 ms.
I/3 H
(P
4 cos 6t (tj I/I2 F
PROBLEM 10.14
^IHF
PROBLEM 10.15
10.16 Find the steady-state value of v using phasors and voltage division.
3 0. I H i n
-A/vv nsw-
1/2 H
9 cos 2
V
t
© io
1/4 F
PROBLEM 10.16
10.17 Find the steady-state value of / if (a) co = 4 rad/s and (b) co = 2 rad/s.Note that
in the latter case the impedance seen at the terminals of the source is purely
resistive.
I H 4 n
10 cos oot
V
© 1/4 F
PROBLEM 10.17
I H I H
-w-
10 cos
V
t
© 3 n
PROBLEM 10.18
_!_^ 1/2 H H
—nm^- I
^rkp —
2 H
4 cos
V
t
© I F
1/4 F
PROBLEM 10.19
I/2(jF
PROBLEM 10.20
I kO 3 kO.
-A/W J VW^
PROBLEM 10.21
PROBLEM 10.22
"g o
2kQ
PROBLEM 10.23
I H
2 cos
V
t
© ^ I F
PROBLEM 10.24
Chap. 10 Sinusoidal Excitation and Phasors 279
4 n h
_j_^
—\W
i
n5W —
20 cos
V
3 t
© 1/3 F
PROBLEM 10.25
10.26 Determine ;'(0) and r(0) in Prob. 10.25 so that the natural component vanishes
and i is simply the forced component.
11
AC STEADY-STATE
ANALYSIS
In the previous chapter we have seen that in the case of circuits with sinusoidal inputs
we may obtain the ac steady-state response by analyzing the corresponding phasor
circuits. The circuits encountered were relatively simple ones which could be analyzed
by the use of voltage-current relations and current and voltage division rules.
It should be clear, because of the close kinship between phasor circuits and resistive
circuits, that we may extend the methods of Chapter 10 to more general circuits using
nodal analysis, loop analysis, Thevenin's and Norton's theorems, superposition, etc.
In this chapter we shall formally consider these more general analysis procedures,
limiting ourselves to obtaining the forced, or ac steady-state, response.
V= ZI
for passive elements is identical in form to Ohm's law, and KVL and KCL hold in
phasor circuits exactly as they did in resistive circuits. Therefore the only difference
in analyzing phasor circuits and resistive circuits is that the excitations and responses
are complex quantities in the former case and real quantities in the latter case. Thus
we may analyze phasor circuits in exactly the same manner in which we analyzed
resistive circuits. Specifically, nodal and mesh, or loop, analysis methods apply. We
shall illustrate nodal analysis in this section and loop analysis in the following section.
To illustrate the nodal method, let us find the ac steady-state voltages v x
and v 2 of
Fig. 11.1. First we shall obtain the phasor by replacing the element values by
circuit
their impedances for co = 2 rad/s and the sources and node voltages by their phasors.
280
Chap. 11 AC Steady-State Analysis 281
This results in the circuit of Fig. 1 1.2(a). Since we are interested in finding V, and V2 ,
the node voltage phasors, we may replace the two sets of parallel impedances by their
equivalent impedances, resulting in the simpler, equivalent circuit of Fig. 1 1.2(b).
(a
2 ° V |
v2
•jin
+ J2
fU
I
5 :&) a 5/0.
Q)
(b)
V V —V
2(V, - 5/0) + -±j- + l
_j{
2
= °
{
v -v,
2 ,
+ J2)/5 1
282 AC Steady-State Analysis Chap. 11
(2+y2)V -yiv 2 1
= io
-/lV 1 + (l-./l)V2 = 5
Solving these equations by determinants, we have
10 -jl
V:
5 1 - j\ 10-75
= 2-;iv
2 + j2 -jl
-jl 1-/1
2+j2 10
V, = yT/-26.6°V
2VT /63.4° V
Therefore the time-domain solutions are
T cos
- 26.6°) V
(2?
the two unknown node voltages v and v + 3000/, as indicated. The phasor circuit in
its simplest form is shown in Fig. 11.4, from which we may observe that only one
nodal equation is needed. Writing KCL at the generalized node, shown dashed, we
have
30001
+ =
^(io 3 )Tf(i-;2)(i03) '
(2-./i)(io 3 )
v+30001
AAAr
3000i 2k.fl
V
4 cos 5000 1 f* J 2kn *»F*
-i" F
T
FIGURE 11.3 Circuit containing a dependent source
V
U 2
vw
k
° /'
'
V V+3000I\
V
\ 3000 I
I
4-
A(IO')
1 = 24 x 10- 3 /53.1° A
= 24 /53.1° mA
Therefore in the time domain, we have
vg = Vm cos oil V
We note first that the op amp and the two 2-kfi resistors constitute a VCVS with
IpF
Vg O A^A
V, - Vm /0 V, - (V/2) ,
V. -V _ n
^
,
(1/VTX10V yT(10 3
) -yl0 6 /co
(V/2) V/2
+ =
2(10 3 )
!
-j\0 6 /co
Eliminating \ x
and solving for V results in
2Vm
[1 -((D 2 /\0 6 )]+j(^f2eo/l0 3 )
In polar form this is
y_ 2Kw /fl
(11.1)
VI + (co/ 1000) 4
where
2~co/1000
9 = -tan" 1
- 2
(11.2)
.1 (co/1000) J
= 2K.
v cos (cor + 0) (11.3)
VI + (co/ 1000)'
We might note in this example that for low frequencies, say < co < 1000, the
amplitude of the output voltage v is relatively large, and for higher frequencies, its
amplitude is relatively small. Thus the circuit of Fig. \. 5 filters out higher frequencies 1
and allows lower frequencies to "pass." Such a circuit is called a filter and will be
considered in more detail in Chapter 15.
-
j iooo/ go ka
I A/2 k A
Vm /Oo —AAA
-j lOOO/co kfl
EXERCISES
11.1.1 Find the forced response v using nodal analysis. Ans. 2 sin 2t V
in
8 cos 2t
A © 4sm 2t
EXERCISE 11.1.1
11.1.2 Find the forced response v. Ans. 4.8 cos (2/ - 53.1°)
3cos2t V
/"T\4sm2t
EXERCISE 11.1.2
which was obtained, using nodal analysis, in the previous section. We shall use the
phasor circuit of Fig. 11.2(b), which is redrawn in Fig. 1.7, with mesh currents I, and
1
286 AC Steady-State Analysis Chap. 11
jl A
b[0_
*fO in
\a A
I, = 2(5 - V,)
or
V, *-* (11.4)
^1,-71(1,-1,) =5
(11.5)
-;l(I 2 -I )-;H + (i^)(I
1 2 2 5) =
I, =6 +jl A
which substituted into (11.4) yields
v = 2-yi v
1
This is the same result that was obtained in the previous section and may be used to
obtain the time-domain voltage v x
.
The same shortcut procedures for writing loop and nodal equations, discussed
in Sees. 4.1 and 4.5 for resistive circuits, apply to phasor circuits. For example, in
Fig. 11.7 if I = —5 is the mesh current in the right mesh in the clockwise direction,
3
(i-yi)i,-(-7i)i 2 = 5
These are equivalent to (1 1.5) and are formed as in the resistive circuit case. That is,
in the first equation the coefficient of the first variable is the sum of the impedances
around the first mesh. The other coefficients are the negatives of the impedances
Chap. 11 AC Steady-State Analysis 287
common mesh and the meshes whose numbers correspond to the currents.
to the first
The right member the sum of the voltage sources in the mesh with polarities consis-
is
tent with the direction of the mesh current. Replacing "first" by "second" applies to
the next equation, and so forth. The dual development, as described in Sec. 4.1, holds
for nodal equations.
As a final example, let us consider the circuit of Fig. 1 1 .8(a), where the response is
the steady-state value of ?•,. The phasor circuit is shown in Fig. 1 1.8(b), with the loop
currents as indicated.
sin 2t A
4/0 + j2a
(b)
V,=./l(4-I)
288 AC Steady-State Analysis Chap. 11
~ 4+
V, = -/3
= 1 /143.1° V
EXERCISES
11.2.1 Find the forced response i in Fig. 11.3 using mesh analysis.
11.2.3 Find the steady-state current i using loop analysis. Ans. cos (t 36.9°) A
en
10 cos +\
t
(
EXERCISE 11.2.3
Because the phasor circuits are exactly like the resistive circuits except for the nature
of the currents, voltages, and impedances, all the network theorems discussed in
Chapter 5 for resistive circuits apply to phasor circuits. In this section we shall illustrate
superposition, Thevenin's theorem, Norton's theorem, and the proportionality prin-
ciple, as applied to phasor circuits.
In the case of superposition, if a phasor circuit has two or more inputs, we may
find the phasor currents or voltages due to each input acting alone (i.e., with the others
dead) and add the individual corresponding time-domain responses to obtain the total.
In the case of a circuit like Fig. 11.1, we may solve the corresponding phasor circuit
of Fig. 1 1.2 by mesh or nodal analysis or by superposition because both sources are
operating at the same frequency, namely to = 2 rad/s. If the sources have different
To illustrate superposition, let us find the forced response i in fig. I 1. 9. There are
two sources, one an ac source with co 2 rad/s and one a dc source with o) 0.
Therefore
/ /,
acting alone. Using phasors, we may find and i 2 by finding their respective phasor
/',
(b). respectively. Figure II. 10(a) is a phasor circuit representing the time-domain
circuit with the current source killed and <x> = 2 rad/s. Figure 1.10(b) is also a phasor 1
3+j2n 2a
-jin in <
D
—
*V <
o
(a )
4Z0 A
(b)
In the latter case, co = 0, the inductances are short circuits (Z L = jO = 0) and the
capacitance is an open circuit (Z r = \/jcoC, which becomes infinite as w— 0). Also,
since the current source is
i
g
= 4 cos (0/ + 0) A
its phasor representation is
I, = 4/01 A
In short, this is just the dc case considered earlier, as is evident from Fig. 11.10(b).
From Fig. 11.10(a) we have
5/0°
I,
3 +J2 + [(1 +y2)(-;l)/(l + J2 - jl)]
T/-8.1
from which
*--(rb)« = -'£ A
from which, since co = 0,
/2 = - A 1
* =h +H
= ^Tcos(2/ - 8.1°) - 1 A
For an example of a circuit with two sinusoidal sources with nonzero frequencies,
the reader is referred to Ex. 11.3.1. The procedure is, of course, exactly the same as
in the preceding example.
In the case of Thevenin's and Norton's theorems the procedure is identical to that
for resistive circuits. The only change is that v oc and iK the time-domain open-circuit ,
(of the dead circuit). There must be only a single frequency present, of course; other-
wise we must use superposition to break the problem up into single-frequency prob-
lems, in each of which Thevenin's or Norton's theorem may be applied.
In general, the Thevenin and Norton equivalent circuits in the frequency domain
are as shown in Fig. 11.11. There is, of course, a close similarity with the resistive case.
As an example, let us use Thevenin's theorem to find the forced response v of Fig.
11.12. We shall find its phasor representation V using the Thevenin equivalent of the
phasor circuit to the left of terminals ci-b. The open-circuit phasor voltage VK is found
from Fig. 1 1.13(a), and the short-circuit phasor current is found from Fig. 11.13(b).
—
6 <b
7 th
(a) (b)
If
a
* o-
+ 2v, A
2 cos
A
3
•0 iF±,
V PC
T
th — ~j (1 1.6)
In Fig. 11.1 3(a), since terminals a-b are open, the current 2/0° flows in the resistor
and the current 2\ flows
x
in the capacitor. Therefore by KVL we have
V« = V, - (-;1)(2V 1 )
where
V, = 2(1) = 2 V
Thus we have
V = 0C 2+ 7-4V
In Fig. 11.13(b), the two nodal equations needed are
V, V,
2V, = 2
1 '
-j\
and
lK = -V, + 2
292 AC Steady-State Analysis Chap. 11
-jio
11
a
/T\
+ A
K
2V,
^ in <
< •
o
(b)
I sc = 3+7lA
The Thevenin impedance, by (11.6), is therefore
The Thevenin equivalent circuit is shown in Fig. 11.14, with the ^-F capacitor, corre-
sponding to —j\ Q, connected to terminals a-b.
It is a simple matter now to see, by voltage division, that
-n (2 + /4)
ui +;i) + (-7ij
4-/2
2 /V/T/-26.6°V
Chap. 11 AC Steady-State Analysis 293
i+ji n
2 + j4
V
6 - j m -p v
As a final topic in this section, let us consider the ladder network of Fig. 11.15
and use the proportionality principle to obtain the steady-state response V. This
requires, as in the resistive case, that we assume V
to have some convenient value like
V= 1 and work backward to find the corresponding \ g Then the correct value of .
V= 1 V
Then from the circuit we have
v y_ 1 + JIA
Continuing, we have
\g = ii 2 + v,=yi +;i=/2v
Therefore V= 1 is the response to \ g =j2. If we multiply this value of \g by 6/J2
to get the correct value of \g , then by the proportionality principle, we must multiply
the assumed response of 1 by the same factor, 6//2, to get the correct value of V.
Therefore we have
V=(£)(1)=-/3V
EXERCISES
11.3.1 Find the forced response /'.
MM6cos(3t-l5°)
EXERCISE 11.3.1
Since phasors are complex numbers, they may be represented by vectors in a plane,
where operations, such as addition of phasors, may be carried out geometrically.
Such a sketch is called a phasor diagram and may be quite helpful in analyzing ac
steady-state circuits.
To illustrate, let us consider the phasor circuit of Fig. 11.16, for which we shall
draw all the voltages and currents on a phasor diagram. To begin with, let us observe
R L
Vo - v,
v
°© Vr -T- c
that the current I is common to all elements and take it as our reference phasor,
denoting it by
I I|/0°
We have taken the angle of I arbitrarily to be zero, since we want I to be our reference.
We may always adjust this assumed value to the true value by the proportionality
principle discussed in the previous section.
The voltage phasors of the circuit are
V R = R\- R\\\
V L = jcoLl = coL\I |/9CT
and
V, = VR + \L + \ c
These are shown in the phasor diagram of Fig. 11.17(a), where it is assumed that
|VJ > |V C |. The cases, |V L |
< |V C and |VJ
|
= |VC |, are shown in Figs. 11.17(b)
and (c), respectively. In all cases the lengths representing the units of current and
voltage are not necessarily the same, so that for clarity I is shown longer than V*.
In case (a) the net reactance is inductive, and the current lags the source voltage
by the angle 6 that can be measured. In (b) the circuit has a net capacitive reactance,
and the current leads the voltage. Finally, in (c) the current and voltage are in phase,
since the inductive and capacitive reactance components exactly cancel each other.
These conclusions follow also from the equation
(11.7)
Z R j[coL - (1/wC)]
\*
vR -v
g
v
c V
Vet Vc t
coL =
coC
or
co = (ll.8)
.s/LC
If the current in Fig. 1 1. 16 is fixed, then the real component of the voltage V^ is
1
l„ Vm (R-jcoL)
R + jcoL R 2 + co L 2 2
Therefore if
I = X .+ jy (11.9)
we have
RV,
x = Re I (11.10)
R2 I co
2
L 2
— m ~coLVm (11.11)
y I I
R2 co
2
L2
The equation of the locus is the equation satisfied by x and y as R varies ; thus we need
to eliminate R between these last two equations.
\ _R_
y coL
from which
coL.x
R
x2 +y = - coL
2
: !
vm y= ( m \ (11.12)
(;•
2coL} \2coL)
which is the equation of a circle with center at [0, — {VJ2coLj\ and radius VjloiL.
The circle (11.12) appears to be the locus, as R phasor I = x
varies, of the jy.
R=
FIGURE 11.20 Locus of the phasor I
298 AC Steady-State Analysis Chap. 11
If Iis as shown in Fig. .20, the current phasor may be resolved into two compo-
1 1
nents, one having amplitude I m cos 6 in phase with the voltage and one with amplitude
l m sin 6, which is 90 out of phase with the voltage. This construction is indicated by
the dashed and dotted vertical line. As we shall see in the next chapter, the in-phase
component of the current is important in calculating the average power delivered by
the source. Thus the phasor diagram gives us a method of seeing at a glance the maxi-
mum in-phase component of current. Evidently this occurs at point a, which corre-
sponds to 6 = 45°. This is the case x = —y, or R = coL.
EXERCISES
11.4.1 Eliminate coL in (11.10) and (11.11) and show that, as coL varies, the locus of
the phasor I = x + j'y is a semicircle.
K.V +
Ans.(,-^y y> = (^)\y<0
2R>
11.4.2 Find coL in Ex. 1.1.4.1 so that Iml has its largest negative value. Also find I for
this case. Ans. R, (VJ^/2 R)l -45°
PROBLEMS
11.1 Solve for the steady-state response / in Fig. 10.19(a) using nodal analysis.
11.2 Solve for the steady-state response i in Fig. 10.20(a) using nodal analysis.
11.5 Find the steady-state value of v if v g = Vm cos cot and find the amplitude of v
for the cases CO= (a) 0, (b) 1000 and (c) 10 6 rad/s. Compare the results for this
circuit with those for Fig. 11.6.
PROBLEM 11.5
Chap. 11 AC Steady-State Analysis 299
—^w^ 2H ia
^vVv-
—
1
nm^-
IH in I H
4 cos
*6
V
I F ^ v
2F^~ in
PROBLEM 11.8
o —— 1*
Z 1
z 3
I
m
z5
+ V
Z2 z4
o -4 l —
PROBLEM 11.9
11.10 Show that the circuit of Prob. 1 1 .8 is a balanced bridge, with the series combina-
tion of 1 Q.and H constituting Z 5
1 in Prob. 1 1 .9. Replace Z 5 by a short, and
show that the same v is obtained as before.
2v, A
PROBLEM 11.11
6 cos 3000 1 V? ±
PROBLEM 11.12
IF 2a
-A/W "
+
4H 4n > v
•g
2
CD r
2D 8
PROBLEM 11.13
4n
e
4 cos 2t V
IH
6 cos 2
V
•6 Ad
2a
2F -
PROBLEM 11.14
"go
PROBLEM 11.15
'g »
, o v
PROBLEM 11.16
•g o o v
PROBLEM 11.17
11.18 Find the forced response r if (a) v g = 2 cos 2/ V and (b) v g = 2 cos / V.
"go
PROBLEM 11.18
PROBLEM 11.19
8a 2H
>A/V 'TJoTT 1
PROBLEM 11.20
PROBLEM 11.21
PROBLEM 11.22
11.23 For the phasor circuit corresponding to Prob. 11.12, replace the part to the
left of terminals a-b by its Thevenin equivalent and find the steady-state value
of/.
11.24 Replace the phasor circuit of the figure for Prob. 11.13, except for the 4-fl
resistor, by its Thevenin equivalent, and find the forced response v.
11.25 Replace the circuit to the left of terminals a-b by its Thevenin equivalent, and
find V.
j3A
-nsw
^-j2a
PROBLEM 11.25
1 1.26 Use the principle of proportionality on the corresponding phasor circuit to find
the steady-state value of v.
PROBLEM 11.26
11.29 Kind 1. the phasor representation of i, using a phasor diagram. Show the
phasors of iR .
c and
f , /,. with the phasor of the source voltage as reference.
PROBLEM 11.29
this case.
In this chapter we shall consider power relationships for networks which are excited
by periodic currents and voltages. We shall concern ourselves primarily with sinu-
soidal currents and voltages since nearly all electrical power is generated in this form.
Instantaneous power, as we now well know, is the rate at which energy
is absorbed
exceeded, the output signal will be distorted. Greatly exceeding this input rating may
even permanently damage the amplifier.
A
more important measure of power, particularly for periodic currents and volt-
ages, that of average power. The average power is equal to the average rate at which
is
Our discussion will begin with a study of average power. We shall then consider
superposition once again and introduce a mathematical measure for characterizing
periodic currents or voltages, known as effective or rms values. We shall then consider
the power factor associated with a load and present a complex power. Finally, we shall
describe the measurement of power.
306
Chap. 12 AC Steady-State Power 307
In linear networks which have inputs that are periodic functions of time, the steady-
state currents and voltages produced are periodic, each having identical periods.
Consider an instantaneous power
p = vi (12. 1)
In this case,
= v(t)i(t)
= p(t) (12.2)
Therefore the instantaneous power is also periodic of period T. That is, p repeats
itself every T seconds.
The fundamental period T of p x
(theminimum time in which p repeats itself) is not
1
T=nT x
(12.3)
= RIl, cos
2
cot
= ^(1 + cos2g>0
Evidently, T, = n/co, and, therefore, T — 27V Thus, for this case, n = 2 in (12.3).
This is illustrated by the graph of/? and i shown in Fig. 12.1(a).
If we now take i — IJA -f cos cot), then
In this case, 7\ = T= Injco, and n = 1 in (12.3). This may be seen also from the
graph of the function in Fig. 12.1(b).
Mathematically, the average value of a periodic function is defined as the time
integral of the function over a complete period, divided by the period. Therefore the
average power P for a periodic instantaneous power/? is given by
t, + T,
P=
Y pdtW (12>4)
\
where r, is arbitrary.
i
i(t),p(t)
(a)
T, --T
i(t),p(t)
(b)
1
p dt (12.5)
ffl7",
P= p dt (12.6)
Therefore we may obtain the average power by integrating over the period off or /.
Chap. 12 AC Steady -State Power 309
Let us now
consider several examples of the average power associated with sinu-
soidal currentsand voltages. A number of very important integrals which often occur
are tabulated in Table 2.1. Verification of these integrals is left as an exercise (see
1
Ex. 12.1.1).
f 2 jt/o
f(0 f(t)dt, co^O
Jo
r. cos (a — P)jco, in = n
First, let us consider the general two-terminal device of Fig. 1 2.3, which is assumed
to be in ac steady state. If, in the frequency domain,
Z = |Z|/0
is the input impedance of the device, then for
two-
V terminal
device
we have
/ = Im cos (cot + - 0) (12.8)
where
|Z|
y , fin >co
-^ cos (cot + 0) cos (co + <f)
-9)dt
\= VJ± = \\\I±
1 = I J<f> - = \l\ /<f,
-
we have, from (12.9),
where
ang V = 0, ang 1 = —
are the angles of the phasors V and I.
Pr = \RJl
It is worth noting at this point that if / = /dc a constant (dc) current, then
, co = =
= 0, and l m = /dc in (12.8). In this special case, we have, by (12.6),
Pr = RIl
energy during part of the period and release it during the other part, so that the aver-
age delivered power is zero.
A very useful alternative form of ( 2.9) may be obtained by recalling that
and therefore
cos 9
ReZ
Substituting this value into (12. 9) and noting that Vm = |Z|/m we have ,
P=£/'ReZ (12.11)
Let us now consider this result if the device is a passive load. We know from the
definition of passivity in (1.7) that the net energy delivered to a passive load is non-
negative. Since the average power is the average rate at which energy is delivered to
a load, it follows that the average power is nonnegative. That is, P > 0. This requires,
by (12.11), that
Re Z{jco) >
or, equivalently,
If 9 = 0. the device is equivalent to a resistance, and if9 = n/2 (or —n/2), the
device is equivalent to an inductance (or capacitance). For — n/2 < 9 < 0, the device
is equivalent to an RC combination, whereas for < 9 < n/2, it is equivalent to an
RL combination.
Finally, if |0| > n/2, then P < 0, and the device is active rather than passive.
In this case the device is delivering power from its terminals and, of course, acts like
a source.
As an example, let us find the power delivered to each element of Fig. 12.4. The
impedance across the source is
Z = 100+./100
= lOOv^T /45° Q
ioon
^
lm
izi ~^T
Therefore, from (12.9), the power delivered to Z is
P= -j(-7j) (100) = 25 W
The power absorbed by the 100-Q resistor is
P fi
= ^(ioo)(i /v^-)* = 25W
2
This power, of course, is equal to that delivered to Z since the inductor absorbs no
power.
The power absorbed by the source is
Pg = -£ffi£-cos0= -25 W
where the minus sign is used because the current is flowing out of the positive terminal
of the source. The source therefore is delivering 25 W to Z. We note that the power
flowing from the source is equal to that absorbed by the load, which illustrates the
principle of conservation of power.
EXERCISES
12.1.1 Verify the integrals of Table 12.1.
12.1.2 For a capacitor of C farads carrying a current / = Im cos cot, verify that the
average power is zero from (12.6). Repeat this for an inductor of L henrys.
12.1.3 Find the average power delivered to a 100-Q resistor carrying a current given
by (a) 10 cos 10/ mA, (b) a square wave for which one cycle consists of 10 mA
1 1
for 1 s followed by —10 mA for 1 s, and (c) a triangular wave for which one
cycle consists of a current that increases linearly from to 10 mA in 1 s.
Ans. (a) 5 mW, (b) 10 mW, (c) 3.33 mW
12.1.4 Find the average power absorbed by each resistor, the capacitor, and the
source. Ans. 20 mW, 10 mW, 0, -30 mW
Chap. 12 AC Steady-State Power 31 3
10 COS lOt I
mF
mA
EXERCISE 12.1.4
12.1.5 lf/i(0 is periodic of period 7", and/2 (0 is periodic of period T2 show that ,
fiU) + fiit) is periodic of period T if positive integers m and exist such that /;
T = mTi = nT2
Extend this result to the function (1 + cos O)?) 2 considered in this section, to
,
In this section we shall consider the power in networks containing two or more
sources, such as the simple circuit of Fig. 12.5. By superposition, we know that
i = U +h
where /, and /
:
are the currents in R due to v gi and vg2 , respectively. The instantaneous
power is
p = R{i + Y {
i2
= Pi + p + 2Ri 2 1
i2
where /?! and p 2 are the instantaneous powers, respectively, due to v gl acting alone
and to vg2 acting alone. In general, 2RiJ 2 0; thus p ^
p + p 2 and superposition ^ x
,
-iripdt
'T
1 f
j (Pi +^2 + 2Ri x
i
2 )dt
P 1
+P + 2R
^T 2 | hi a2
/,/ A
JO
where Pj and P2 are the average powers from v gl and v g2 , respectively, acting alone.
(We are assuming, of course, that each component of p is periodic of period T.)
Superposition for average power applies when
P= P X
+P 2 (12.12)
The most important case in which this equation is satisfied is when i(t) is composed
of sinusoidal components of different frequencies. Suppose, for instance, that
and
where w and « are positive integers. Therefore if co is a number such that T = 2n/co,
then co l = mco and co 2 = nco. In this case the integral in (12.13) becomes, using Table
12.1,
rT p2 n/oi
ij 2 dt = Im Jm2 cos (mcot + $.) cos (ncot -f $2) rfr
Jo Jo
_ 7 wl /m2 ^cos(0 1
- 2) .
m= n
CO
= 0, /w ^ «
Chap. 12 AC Steady-State Power 315
It can be shown that superposition of average power holds for sinusoids whose
frequencies are not integral multiples of some frequency co, provided we generalize
the definition of average power to
= r —]
P pdt
—
lim
T Jo
This generalization applies to the periodic case just considered as well as to the case
i = *! + i2 , where
f, = cos t
U = cos nt
In this case i is not even periodic (the ratio coJco 2 = \jn is not a rational number
mjn), but
lim —T dt =
T —oo
r«
| /,/,.
I, == A
1/60°
I2 = -0.5 A
Therefore
1 = 1, +i =yo.866A
2
p= ^-(100)(0.866)
2
= 37.5 W
Let us now repeat the above example with v g2 = 50 V. Since v gl and vg2 are sinu-
soids with co l
= 377 and co 2 = rad/s, respectively, superposition for the average
power is applicable. Proceeding as before, we find
I2 = -0.5 for co =
.
P
tl W "'ml
2
= 50
Pi = RIli = 25 W
and the average power is
p = p + p2 = i
75 W
This example illustrates a very important case for electronic amplifiers with
sinusoidal inputs. These amplifiers contain dcpower supplies that produce dc currents
which provide the energy for the amplified ac signals. Thus superposition is very
useful in finding the average power associated with each frequency, including co = 0.
Extending the above procedure to a periodic current which is the sum of N + 1
sinusoids of different frequencies,
The first term, RIl,,, in which the factor £ is missing, is the special case of zero
frequency and must be considered separately, as in Sec. 12.1. That is, by (12.6),
and
T
z
RI dc dt
EXERCISES
12.2.1 Find the average power delivered to the resistor in Fig. 12.5 if R= 10 Q. and
(a) v = 10 cos 100/ and v g2 = 20 cos (100/ + 30°) V.
>
(b) V tl
<) = 100 cos (/ + 60°) and v gl = 50 cos (2/ - 30°) V.
(c) Vy t i = 50 cos (/ + 45°) and v g2 = 100 sin (/ + 30°) V.
(d) V gl = 20 cos (/ + 25°) and v g2 = 30 sin (5/ - 50°) V.
Arts, (a) 7.68 W, (b) 625 W, (c) 754.4 W, (d) 65 W
1
12.2.2 Miul the average power absorbed by each resistor and each source.
Am. 3.5 W, 3.5 W, -5 W, -2 W
v^W ? >
1 kfl
100 V ~
T l>4F
1 1 *
EXERCISE 12.2.2
RIL Ri 1 dt
/rn
M" dt (12.16)
^ J
COS
2
(cot + 0) (ft
Thus a sinusoidal current having an amplitude / m delivers the same average power
to a resistance R as does a dc current which isequal to IJ^J~1 We also see that the .
rms current is independent of the frequency co or the phase <j) of the current Similarly, /'.
Substituting these values into the important power relations of (12.9) and (12.1 1) for
the two-terminal network, we have
and
P= I;ms Re Z (12.18)
\ SfcCs or>\*)
In practice, rms values are usually used in the fields of power generation and^
For instance, the nominal 15-V ac power which is commonly used for
distribution. 1
household appliances is an rms value. Thus the power supplied to our homes is
provided by a 60-Hz voltage having a maximum value of 115,^/T «= 163 V. On the
other hand, maximum values are more commonly used in electronics and com-
munications.
Finally, let us consider the rms value of the current in (12.14), which is made up
of sinusoids of different frequencies. In terms of rms currents, (12.15) becomes
Since P= 2
i?/ r ms , we see that the rms value of a sinusoidal current consisting of
different frequencies is
"rms — -V 'dc i
1 1 , m, + * 2,*, ~1~ • • \ * f/, m .
Similarly,
These results are particularly important in the study of noise in electrical networks,
a subject of later courses.
Chap. 12 AC Steady -State Power 319
EXERCISES
12.3.1 Find the rms value of a periodic current for which one period is defined by
(a) i /. 0< /</,
= 0, /, < <t (T=t z
t 2 ).
EXERCISE 12.3.3
"^
P = ^rms'rms COS d
^^
The power is thus equal to the product of the rms voltage, the rms current, and the
cosine of the angle between the voltage and current phasors. In practice, the rms
current and voltage are easily measured and their product, Krms /rms , is called the
apparent power. The apparent power is usually referred to in terms of its units, volt-
amperes (VA) or kilovoltamperes (kVA), in order to avoid confusing it with the unit
of average power, the watt. It is clear thai the average power can never be greater than
the apparent pow er.
The ratio of the average power to the apparent power is defined as the power
factor. Thus if we denote the power factor by pf, then in the sinusoidal case
Pf= V I
cos 9 (12.21)
320 AC Steady-State Power Chap. 12
(pf = 1) exist for loads which contain inductors and capacitors if the reac-
can also
tances of these elements are such that they cancel one another. Adjusting the reactance
of loads so as to approximate this condition is very important in electrical power
systems, as we shall see shortly.
In a purely reactive load, = ±90°, pf — 0, and the average power is zero.
In this case, the equivalent load is an inductance (9 = 90°) or a capacitance (9 =
—90°), and the current and voltage differ in phase by 90°.
A load for which —90° < 9 < is equivalent to an RC combination, whereas
one having < 9 < 90° is an equivalent RL = cos — 9),
combination. Since cos 9 (
it is evident that the/?/ for an RC load having = — where < 0, < 90°, is equal 15
60 Hz has Z = 100 +;37.7 = 106.9 /20.66° fl and has a pf of cos 20.66° = 0.936
lagging.
In practice, the power factor of a load is very important. In industrial applications,
for instance, loads may require thousands of watts to operate, and the power factor
greatly affects the electric bill. Suppose, for example, a mill consumes 100 kW from
a 220-V rms line. At a/?/ of 0.85 lagging, we see that the rms current into the mill is
P 105
534.8 A
rms
K ms pf (220)(0.85)
Now suppose that the pf by some means is increased to 0.95 lagging. Then
10 s
/
rms
= — 478 S A
(220)(0.95)
Comparing the latter case with the former, we see that 7 rms was reduced by 56.3 A
(10.5%). Therefore the generating station must generate a larger current in the case
of the lower/;/. Since the transmission lines supplying the power have resistance, the
generator must produce a larger average power to supply the 100 kW to the load.
Chap. 12 AC Steady-State Power 321
If the resistance is 0.1 Q. for instance, then the power generated by the source must be
/> =10 + 5
0.1/?n
Therefore we find
which requires that the power station produce 5.7 kW (4.64 "J more power to supply
the lower/'/' load. It is power companies encourage apf exceeding
for this reason that
say 0.9 and impose a penalty on large industrial users who do not comply.
Let us now consider a method of correcting the power factor of a load having an
impedance
Z = R + jX
not affected. The current I, supplied by the generator, however, does change.
z T — LJ z. z R + ,X
— zz,
Mwp
z,
In general, we select Z so that (1) Zj absorbs zero average power, and (2) Z T satisfies
x
the desired power factor pf = PF. The first condition requires that Z, be purely
reactive. That is,
j*i
/Im Z
cos tan
,
PF
VRe Z T )]
322 AC Steady-State Power Chap. 12
7?2 _L V2
v (12.22)
1
k iRtun {cos- l
PF)-X
where &i = 1 if the power factor is lagging upon correction and k^ =— 1 if the power
factor is leading upon correction. In each case, tan (cos -1 PF) is taken as positive.
As an example of the application of (12.22), let us change the power factor for the
Z = 100+y'100= 141.4/45°
Since X < x
0, the reactance is a capacitance C = —\/coX = l
33.6 ^F. The load
impedance now becomes
7 _ (100 +7l00)(-/297.92) _
P= cos(18 2 -
°
)
= 25W
2TTW)
which is the same as that delivered to Z in Fig. 12.4. The current, however, is
-0.5 A
2
We see, therefore, that the current has been reduced by 0.128 A or 25.6%.
A
EXERCISES
12.4.1 Find the apparent power for (a) a load that requires 30 Arms from
a 230-V mis line and (b) a load consisting of a 100-Q resistor in parallel with
a 25-//F capacitor connected to a 1 10-V rms 60-Hz source.
Ans. (a)6.9kVA, (b) 166.3 V
12.4.2 Find the power factor for (a) a load consisting of a series connection of a 100-Q
and a 20-//F capacitor operating at 60 Hz, (b) one that is capacitive
resistor
and requires 50 A rms and 5 kW at 10 V rms, and (c) one consisting of
1
12.4.3 A load consists of a 200-Q resistor in series with a 0.1-H inductor. Find the
parallel capacitance necessary to adjust to (a) a p/of 0.95 lagging and (b) a pf
of 0.95 leading if CO = lOOOrad/s. Ans. (a) 0.685 p.¥, (b) 3.31 //F
We shall now introduce a complex power in the ac steady state which is very useful
for determining and correcting power factors associated with interconnected loads.
Let us begin by defining rms phasors for general sinusoidal voltages and currents.
The phasor representations for (1 2.7) and (12.8) are
V = Vm e*
\ — i P m-e)
V = V
V rms =
* /~0~ rms e'* 1
(12.23)
I
*rms
= r~*r
— J/ rms c
ptt4>-&)
Let us now consider the average power given in (12.17). Using Euler's formula,
we may write
P = yrm J rm s
cos 6 - Re( VTm ,Irmi e»)
V I* = V 1 e'
e
324 AC Steady-State Power Chap. 12
*
where I r ms is the complex conjugate of Irms . Thus
and the product V rms I*ms is a complex power whose real part is the average power.
Denoting this complex power by S, we have
where Q is the reactive power. Dimensionally, P and Q have the same units; however,
the unit of Q is defined as the var (voltampere reactive) to distinguish it from the watt.
The magnitude of the complex power is
I
SI
kJ
I
= IV
I
I*
T rms*rms I
I
= IV
I
T rras III*
I
J rras
I
I
I
= Vrms'rms
1 r
Q=y jrms
ImZ
\L ' rras 1 i
7 i
Q= 2
/ r ms Im Z
= ^tw (12.27)
A phasor diagram of V rms and I rms is shown in Fig. 12.7. We see that the phasor
current can be resolved into the two components I!ms cos and /rms sin 6. The com-
ponent 7rms cos 9 is in phase with V rms and it produces the real power P. In contrast,
,
7rms sin 9 is 90° out of phase with V rras and it causes the reactive power Q. Since 7rms
,
sin 9 is 90° out of phase with V rms it is called the quadrature component of I rms As a
, .
that of Fig. 12.8. It is evident that for an inductive load (lagging pf), < 9 90°, <
Q is positive, and S lies in the first quadrant. For a capacitive load (leading pf),
— 90° 9 < 0, Q is negative, and S lies in the fourth quadrant. A load having
<
a unity power factor requires that Q — since 9 = 0. In general, we see that
9 = tan (12.28)
'(f)
Chap. 12 AC Steady-State Power 325
Re K Re
Let us now consider the complex power associated with a load consisting of two
impedances Z and Z 2 as shown in Fig. 12.9. The complex power delivered to the
l
,
combined impedances is
V rms*rms
T I* V rms\(Ix
T
1 rms l
V rms*T*
T
1 rr
i_
1
v rms*2
*
I* rms
Therefore the complex power delivered by the source to the interconnected loads is
the sum of that delivered to each individual load, and the complex power is thus con-
served. This statement no matter how many individual loads there are or how
is true
they are interconnected, because it depends only on Kirchhoff's laws and the defini-
S = P+jQ
Connecting a pure reactance Z, in parallel with Z results in a complex power to Z, of
S, =JQ t
l,„.
Therefore, from conservation of complex power, for the composite load we have
= P+KQ+Qr)
It is evident that the addition of Z x
does not affect the average power P delivered to
the load. It does, however, affect the net reactive power. We can therefore select Q 1
to obtain our desired power factor, which, of course, is usually raised. This causes
a reduction in the current required to produce P, as discussed previously.
Let us again consider the circuit of Fig. 12.4 and change the power factor to
PF = 0.95 lagging. The complex power of the uncorrected load is
c
*-*
__ v rms'rras
T T* — P+jQ = 25+y'25
since
V rms
T = 70.7 V
—_ V rms
T
= 0.3535(1 -yi)A
-*rms
Z
From (12.28), we see that QT = Q + Q l
must satisfy
= tan-(^)
Therefore
cos^ = PF=cos(tan-'^
and
QT = P tan (cos' 1
PF)
= 25 tan 18.2° = 8.22 vars
The required Q x
is
X = x
^W = -297.9 Q
pfi = 0.9 lagging and Z 2 a 5-kW load with pf2 = 0.95 leading. For Z, we have
s, =P +jQ i i
Chap. 12 AC Steady -State Power 327
where P =
i
10 4 W, 0, = cos" 1
S2 = P2 + ;&
where P =
2
5 x 10 3 W, 9 2 = -cos" pf2 1
= -18.2°, and
ST = P + P2 +j(Q + Q
1 1 2)
= 1.5 x 10 4
-fy'3200
°
= an 04
* '
-'(rOT) = 12 -
and
EXERCISES
12.5.1 Find the complex power delivered to a load which has a 0.91 lagging power
factor and (a) absorbs 1 kW, (b) 1 kvar, and (c) 1 kVA.
Arts, (a) 1099 /24.5° (b) 2412 /24.5°
, ,
(c) 1000/24.5°
12.5.2 Find the impedance of the loads in Ex. 12.5.1 if Vtms = 115 V.
Ans. (a) 12.03 /24.5° (b) 5.48 /24.5°
, ,
(c) 13.23 /24.5°
12.5.3 Repeat Ex. 12.4.2 (c) using the concept of complex power.
A wattmeter is a device which measures the average power that is delivered to a load.
It contains a rotating high-resistance voltage, or potential, coil, connected in parallel
with the load and a fixed low-resistance current coil which is connected in series with
the load.The device has four terminals, a pair to accommodate each coil. A typical
connection is shown in Fig. 12.10. We see that the current coil responds to the load
328 AC Steady-State Power Chap. 12
CURRENT
COIL
VOLTAGE
COIL
current, whereas the voltage coil responds to the load voltage. For frequencies above
a few hertz, the meter movement responds to the average power. Ideally, the voltage
across the current coil and the current in the voltage coil are both zero, so that the
presence of the meter does not influence the power it is measuring.
One terminal on each of the coils is marked ± so that if the current enters the ±
terminal of the current coil and the ± terminal of the voltage coil is positive with
respect to its other terminal, then the meter gives a positive, or upscale, reading. In
Fig. 12.10 this corresponds to the load absorbing power. If the terminal connections
of either the current coil or voltage coil (but not both) are reversed, a negative, or
downscale, reading is indicated. Most meters cannot read downscale —the pointer
simply rests on the downscale stop. Thus such a reading requires reversing the connec-
tions of one of the coils, usually the voltage coil. Reversing the connections of both
coils does not affect the reading.
The wattmeter of Fig. 12.10, represented by the rectangle and the two coils, is
connected to read
P = |V| • |I|cos0
where V and I are as indicated and is the angle between V and I (or, equivalently,
the angle of the load impedance). This, of course, power delivered to the load.
is the
Other types of meters are available for measuring the apparent power and the
reactive power. An apparent power or VA meter simply measures the product of the
rms current and rms voltage. The varmeter, on the other hand, measures the reactive
power.
EXERCISE
12.6.1 Determine the power reading of each wattmeter after assigning the terminal
markings required for a positive reading. Ans. 87.5 W, 9.375 W, W
Chap. 12 AC Steady-State Power 329
ioon
—v>
6n;
ioo rnrn
100
n
©J I A
V rms
Q)
EXERCISE 12.6.1
PROBLEMS
12.1 One cycle of a periodic current is given by
12.2 Determine the average power for p(t) of Fig. 12.1 (b).
->Wv-
ion
10 COS 5t
V
© 0.5 H 0.1 F _~
PROBLEM 12.3
12.4 Repeat Prob. 12.3 with the Jg-F capacitor replaced by a current-controlled
voltage source of value v = 10/ V.
and (b) show that the maximum average power is delivered to a load R L when
330 AC Steady-State Power Chap. 12
~ 0.25 F
2 COS 2t
I n
PROBLEM 12.6
o
12.7 Repeat Prob. 12.6 with the 2-A source replaced by a source having i = 2 cos A t
12.8 We have defined the average power for an instantaneous power p(t), which is
P= lim
1
Pit) dt
Show for v and i of (12.7)— (12.8) that this definition yields the same result as
(12.6).
/, = I lm cos co x
t
'2 = hm cos (o 2 t
is the current flowing in a resistor R. Using the definition of Prob. 12.8, show
for cx>! =£ co 2 that
P= /», + P2
where P, and P2 are the average powers associated with /', and i2 , respectively,
acting alone. Note that this includes the case of pit) being nonperiodic.
12.10 Find the rms value of the voltage (a) v = 100 cos (lOOt + 35°) V and (b)
v = 8 cos It + 4 cos It + b^TL cos (3t - 45°).
12.11 Determine the rms value of a periodic current for which one cycle is given by
12.15 Determine the power factor for the circuit. What parallel reactance will change
the power factor to 0.95 lagging?
30X1
PROBLEM 12.15
12.16 Determine the power factor as seen from the terminals of the independent
voltage source.
I .a
us Z0( ,21 2]
V rms
PROBLEM 12.16
12.17 Determine the reactive element needed to be added in parallel to the source to
power factor of Prob. 12.13 to 0.95 lagging. Assume a frequency of
raise the
100 Hz.
12.18 A load Z L requires 10 kW at a power factor of 0.7 lagging for a load voltage of
230 V rms at 60 Hz. The transmission line connecting the generator to the load
has a series impedance of 0.1 +/0.4D. Determine the generator voltage and
power factor.
12.19 A 230-V rms 60-Hz load draws 500 kW and 400 kvars. Compute the parallel
capacitance necessary to give a power factor of (a) 1.0, (b) 0.9 lagging, and (c)
0.9 leading.
12.20 Two loads in parallel draw a total of 3 kW at a 0.9 lagging pf from a 115-V
rms 60-Hz line. One load is known to absorb 1000 W at a 0.8 lagging power
factor. Find (a) the power factor of the second load and (b) the parallel reactive
element value necessary to correct the power factor to 0.95 lagging for the
combined load.
12.21 A load Z L requiring 5 kW at a 0.8 lagging power factor is supplied from two
60-Hz generators, as shown. Generator Vj has a terminal voltage of 230 V
rms and delivers 2.5 kW at a 0.8 lagging power factor through a transmission
332 AC Steady -State Power Chap. 12
ing V2 has an impedance Z 2 = 0.8 + /1 Q. Find (a) the load voltage V L (b) ,
z, 7
"~z
T
z L vL
O
PROBLEM 12.21
13
THREE-PHASE CIRCUITS
As we have already noted, one very important use of ac steady-state analysis is its
application to power systems, most of which are alternating current systems. One
principal reason for this is that it is economically feasible to transmit power over long
distances only if the voltages involved are very high, and it is easier to raise and lower
voltages in ac systems than in dc systems. Alternating voltage can be stepped up for
transmission and stepped down for distribution with transformers, as we shall see in
Chapter 16. Transformers have no moving parts and are relatively simple to construct,
whereas with the present technology, rotating machines are generally needed to raise
and lower dc voltages.
Also, for reasons of economics and performance, almost all electric power is
produced by polyphase sources (those generating voltages with more than one phase).
In a single-phase circuit, the instantaneous power delivered to a load is pulsating, even
if the current and voltage are in phase. A polyphase system, on the other hand, is
all the power produced in the world is polyphase power at 50 or 60 Hz. In the United
333
334 Three-Phase Circuits Chap. 13
with a different phase. If the three currents drawn from the sources also constitute
a balanced set, then the system is said to be a balanced three-phase system. This is
the case to which we shall restrict ourselves, for the most part.
Before proceeding to the three-phase case, let us digress in this section to establish
our notation and consider an example of a single-phase system that is in common
household use. This example will also serve to give us some practice with a single-
phase system, with which we are already familiar and which will serve as an introduc-
tion to the three-phase systems to be considered next.
In this chapter we shall find extremely useful the double-subscript notation
introduced in Chapter 2 for voltages. In the case of phasors, the notation is \ ab for the
voltage of point a with respect to point b. We shall also use a double-subscript notation
for current, taking, for example, l ab as the current flowing in the direct path from point
a to point b. These quantities are illustrated in Fig. 13.1, where the direct path from a
to b is distinguished from the alternative path from a to b through c.
V Qb
Because of the simpler expressions for average power that result, we shall use rms
values of voltage and current throughout this chapter. (These are also the values read
by most meters.) That is, if
V = |V|/0° Vrms
(13.1)
1 = |I|/-fl Arms
Z = |Z|/0 Q (13.2)
1
P = |V| •
1
1 1 cos <9
1
2
ReZW (13.3)
v = «/2 V | |
cos cot V
i = »/2 1
1 1 cos (cot — 0) A
The use of double subscripts makes it easier to handle phasors both analytically
'
Van : |0 °
V rms
£© V Qb
-0-
°
/„„= l00 /- |20
V rms
(o) (b)
This is evident without referring to a circuit since by KVL the voltage between two
points a and b is the same regardless of the path, which in this case is the path a,n,b.
Also, since \ nb = —\ hn , we have
\ ab = 100.^3730° V rms
A single-phase, three-wire source, as shown in Fig. 13.3, is one having three output
terminals a, b, and a neutral terminal n, for which the terminal voltages are equal. That
is,
V =V v, (13.4)
336 Three-Phase Circuits Chap. 13
This is a common
arrangement in a normal household supplied with both 115 V and
230Vrms, since if |V J = |V,|= 115 V, then |V
fl
= |2V,| = 230 V. fl6 |
Let us now consider the sourceof Fig. 13.3 loaded with two identical loads, both
having an impedance Z,, as shown in Fig. 13.4. The currents in the lines aA and bB
are
aA "
z, z,
and
l„B
z, 'z,
a a
Von© 2 I
Vnb(^ 1
U = -Q.A + 1„b) =
Thus the neutral could be removed without changing any current or voltage in the
system.
If the lines aA and bB are not perfect conductors but have equal impedances Z 2 ,
then I nN is still zero because we may simply add the series impedances Z, and Z 2 and
have essentially the same situation as in Fig. 13.4. Indeed, in the more general case
shown in Fig. 13.5, the neutral current I nN is still zero. This may be seen by writing the
two mesh equations
(Z, +Z +Z 2 3
)(I a/1 + l bB ) +Z 3 (I a ^ + i bB ) = o
Chap. 13 Three-Phase Circuits 337
Vi
9 O z3 N 1
".© ' hP
or
Ka + hB = o (13.5)
Since by KCL the left member of the last equation is —l nN , the neutral current is zero.
This is, of course, a consequence of the symmetry of Fig. 13.5.
If the symmetry of Fig. 13.5 is destroyed by having unequal loads at terminals
A-N and N-B or unequal line impedances in lines aA and bB, then there will be a
neutral current. For example, let us consider the situation in Fig. 13.6, which has two
loads operating at approximately 115 V and one at approximately 230 V. The mesh
equations are
I A
ion
jioo
b B
EXERCISES
13.1.1 Derive (13.5) by superposition applied to Fig. 13.5.
and c and a neutral terminal n. In this case, the source is said to be Y-connected
(connected in a Y, as shown). An equivalent representation is that of Fig. 13.7(b),
which is somewhat easier to draw.
The voltages \ an Y bn
, , and V c„
between the line terminals and the neutral terminal
are called phase voltages and in the cases we shall consider are given by
y an = vP iw
\ bn = K. /-12Q (13.6)
v c„
= r,/120°
or
v an = v,/or
v,l<r
v 6n = VJ120° (13.7)
v „ = VJ-120
c
In both cases, each phase voltage has the same rms magnitude Vp and , the phases are
Chap. 13 Three-Phase Circuits 339
(o) (b)
displaced 120°, with \ an arbitrarily selected as the reference phasor. Such a set of
voltages is called a balanced set and is characterized by
(13.8)
(o) (b)
(13.8) holds. Evidently, the only difference between the positive and negative sequence
is the arbitrary choice of the terminal labels, a, b, and c. Thus without loss in generality
we shall consider only the positive sequence.
By (13.6), the voltages in the abc sequence may each be related to V a „.
The relation-
ships, which will be useful later, are
(13.9)
V en = V an /120°
T — I
l
The line-to-line voltages, or simply line voltages, in Fig. 13.7 are Y ab Ybc
, , and V ca ,
Y ab = V +
a„ Y„ b
= v,l& - vp /-\20°
= J^v p
\w_
In like manner,
y„ c = */TVP l- -90°
yca = «/?v,i- -210°
Vt = */JV. (13.10)
and thus
V ab =VL /W, Vbc = VJ-9Q° , V ca = VL /-2\0° (13.11)
These results also may be obtained graphically from the phasor diagram shown in
Fig. 13.9.
Let us now consider the system of Fig. 13.10, which is a balanced Y-Y, three-phase,
four-wire system, if the source voltages are given by (13.6). The term Y-Y applies since
both the source and the load are Y-connected. The system is said to be balanced since
the source voltages constitute a balanced set and the load is balanced (each phase
impedance is equal — in this case to Z p The
). fourth wire is the neutral line n-N, which
may be omitted to form a three-phase, three-wire system.
The line currents of Fig. 13.10 are evidently
T vY an
V,„ V a „/-120°
bB = W-120° (13.12)
zP z,
V„„ V an /120°
cC ~ z, "
Z,
WI20°
Chap. 13 Three-Phase Circuits 341
O
O^
'bn
The last two results are a consequence of (13.9) and show that the line currents also
form a balanced set. Therefore their sum is
P. = Vj p cos 6
(13.14)
= I, Re Z p
2
P=3PP
The angle 9 of the phase impedance is thus the power factor angle of the three-phase
load as well as that of a single phase.
Suppose now that an impedance Z L is inserted in each of the lines aA, bB, and cC
and that an impedance Z N not necessarily equal to Z L is inserted in line nN. In other
, ,
words, the lines are not to be perfect conductors but are to contain impedances.
Evidently, the line impedances, except for ZN , are in series with the phase impedances,
and the two sets of impedances may be combined to form perfect conducting lines aA,
bB, and cC with a load impedance Z p + Z L in each phase. Therefore, except for Z N in
the neutral, the equivalent system has perfect conducting lines and a balanced load. If
the impedance Z N were not present in the neutral, it would be a perfect conductor,
and the system would be balanced as in Fig. 13.10. In this case, as we have seen,
points n and N are at the same potential, and there is no neutral current. Thus it does
not matter what is in the neutral line. It may be a short circuit or an open-circuit or
contain an impedance such as Z N and still no neutral current would flow and no
,
voltage would appear across nN. Obviously, then, the presence of equal line imped-
ances in aA, bB, and cC and an impedance in the neutral does not change the fact that
the line currents form a balanced set.
As an example, let us find the line currents in Fig. 13.11. We may combine the 1-Q
line impedance and (3 + j3)-Q phase impedance to obtain
Zp = 4+j3 = 5 /36.9° Q
as the effective phase load. Since by the foregoing discussion there is no neutral
current, we have
10 °/0° - 0/ , o
I fiQ A
The currents form a balanced, positive sequence set, so that we also have
100/0° V rms
o
in
O- 3+j3a
100
-Q
/-1
—
20° Vrms
w 3+j3n
3n
100 /120°
-0 — Vrms
S
I
w\a 3+j3a
This example was solved on a "per-phase" basis. Since the impedance in the
neutral is immaterial in a balanced Y-Y system, we may imagine the neutral line to be a
short circuit. We may do this if it contains an impedance or even if the neutral wire is
not present (a three-wire system). We may then look at only one phase, say "phase A"
consisting of the source \ an in series with ZL and Zp , as shown in Fig. 13. 12. (The line
nN is replaced by a short circuit.) The line current \ aA the , phase voltage l aA Zp , and the
voltage drop in the line I^Z^ may all be found from this single-phase analysis. The
other voltages and currents in the system may be found similarly, or from the previous
results, since the system is balanced.
a A
zL
Zp
p n N
1
o o
VJB cos 9
we have
300= 0W) 7,(0.9)
° 5
Therefore since for a Y-connected load the phase current is also the line current, we
have
| Z '' I
Vp 200/yT _ ^
1
/, 3 s A3"/(2)(0.9)
Zp = 40 /25.84° Q
If the load is unbalanced but there is a neutral wire which is a perfect conductor,
then we may still use the per-phase method of solution for each phase. However, if
this is not the case, this shortcut method does not apply. There is a very useful method
employing so-called symmetrical components which is applicable to unbalanced systems
and which the reader may encounter in a course on power systems. It is important to
note, in any case, that a three-phase circuit is still a circuit and, balanced or unbalanced,
may always be solved by general analysis procedures.
EXERCISES
j Z Z
t 2
p
z, + z 2
A o
B o
C o
EXERCISE 13.2.3
shown in Fig. 13.13(a), in a way that resembles a A, and in an equivalent way in Fig.
1 3. 1 3(b). If the source is Y- or A-connected, then the system is a Y-A or a A-A system.
An advantage of a A-connected load over a Y-connected load is that loads may be
added or removed more readily on a single phase of a A, since the loads are connected
directly across the lines. This may not be possible in the Y connection, since the
neutral may not be accessible. Also, for a given power delivered to the load the phase
currents in a A are smaller than those in a Y. On the other hand, the A phase voltages
are higher than those of the Y connection. Sources are rarely A-connected, because if
the voltages are not perfectly balanced, there will be a net voltage, and consequently a
circulating current, around the delta. This, of course, causes undesirable heating
effects in the generating machinery. Also, the phase voltages are lower in the
Y-connected generator, and thus less insulation is required. Obviously, systems with
A-connected loads are three-wire systems, since there is no neutral connection.
From Fig. 13.13 we see that in the case of a A-connected load the line voltages are
the same as the phase voltages. Therefore if the line voltages are given by (13.11), as
before, then the phase voltages are
a o a o-
bo- , B Zp
(o) (b)
-^ = / p /30° - 9
IJ -90° - 9 (13.17)
Ica = ^= I J-2\0°-9
where
Vl
(13.18)
*~aA — *AB *C
ia„ = yr j p /-e
i bB = A/jij-i2o°-e
l cC = /Jl /-240°-9
/S p
Evidently the relation between the line and phase current magnitudes in the
A case is
Chap. 13 Three-Phase Circuits 347
Thus the currents and voltages are balanced sets, as expected. The relations between
line and phase currents for the A-connected load are summed up in the phasor diagram
of Fie. 13.14.
Vab
As an example of a three-phase circuit with a A-connected load, let us find the line
current I L in Fig. 1 3. 1 3 if the line voltage is 250 V rms and the load draws l .5 kW at a
lagging power factor of 0.8. For one phase, Pp = '
5
3
00 = 500 W, and thus
500 = 250/(0.8)
or
Ip = 2.5 A rms
Therefore we have
IL = J3 ip = 4.33 A rms
Finally, in this section, let us derive a formula for the power delivered to a balanced
three-phase load with a power factor angle 6. Whether the load is Y-connected or
A-connected, we have
P=3Pp = 3Vp Ip cos6
348 Three-Phase Circuits Chap. 13
/>=3-£=cos0
or
P = +/TVJL cos6 (13.21)
IL = 4.33 A rms
EXERCISES
13.3.1 Solve Ex. 13.2.2 if the source and load are unchanged except that the load is
A-connected. [Suggestion: Note that in (13.15), (13.17), and (13.20) 30° must
VL = 200 V rms at the load terminals. Find the total power delivered to the
load. Ans. 19.2 kW
13.3.3 If the lines in Ex. 13.3.2 each have a resistance of 0.1 D, find the power lost in
the lines. Ans. 1.44 kW
A B
1 !* 1
I ;
(a) (b)
Y ab Y bc
, , and Y fa of the A in terms of Y a Y b and Y
, ,
c
of the Y so that the A connection
is equivalent to the Y connection at the terminals A, B, and C. That is, if the Y is
replaced by the A, the same node voltages V,,, YB , and Vc will appear, and the same
currents I, and I 2 will flow. Conversely, a A-Y transformation is an expression of the Y
parameters in terms of the A parameters.
Let us begin by writing nodal equations for both circuits. If node C is taken as
reference, in the case of the Y network we have
Y.V^ -YY a fl
Y VB
6
Y,V = I2
Y V^ - Y.V,, + (Y
a a + Y + Y )VD
b C
Solving for YD in the third equation and substituting its value into the first two equa-
350 Three-Phase Circuits Chap. 13
/ Y a Y, + Y a Y c
\ / Y a Y, \ _
\\. + \b + \ c r A
\\ a +\ +\ b c r B ~ li
(13.22)
(
Ya Y 6 W / Y a Y, + Y,Y c \ _
+ Y JYA - Y \ B =
(Y fli ab I,
-Y V, + (Y + Y )\ B =
flfc a6 bc I2
Equating coefficients of like terms in these equations and (13.22) and solving for the
admittances of the A circuit, we have the Y-A transformation
v*ab YaY 6
Y a + Y4 + Y c
v*bc Y4 Y C
(13.23)
Y a + Y, + Y c
Y1 ca Y Ya c
Ya + Y + Y b c
words, as follows:
Z„6 +z +z ic ca
z =
z z fcc lb
fc
z fl6 + z, c + Zca
z fa z be
Z a6 + Z bc + Z ca
where the Z's are the reciprocals of the Y's of Fig. 13.16. The rule is as follows:
(By adjacent here we mean "on each side of and terminating on the same node as." For
example, in the superimposed drawing of the Y and A, Z lies between Z ah and Zca and
all three have a common terminal A. Thus Z ab and Z ca are adjacent arms of Z a .)
As an example, let us find the input impedance Z of Fig. 3. 7(a). This is a problem
1 1
which would have required us to write loop or nodal equations in the past, because we
cannot simplify the circuit by combining series and/or parallel impedances. Replacing
the l-. I-. and ^-Q resistors, which constitute a Y, by their equivalent A, as shown in
Fig. 13.17(b). however, enables us to solve the problem readily.
60 6H
4n 4fl
1(2)
1+2 + 3
Y bc 2(3)
= 1
1+2 + 3
Y, „ = 3(1)
1+2 + 3
3 a i a 2fi
Thus Fig. 13.17(b) may be simplified by combining parallel and series resistors to
obtain
Z=3 lQ g
since Ya Y ,
4 , and Y c
are all equal to Y^,, the reciprocal of Zy , then the equivalent
A-connected load is also balanced because
YM ab = Y be = Y ca = t\/ Y
Yy = — 2
-5
Thus if Zd is the phase impedance of the equivalent balanced A-connected load, then
Z, = 3Z„ (13.25)
EXERCISES
13.4.1 Find the input impedance seen by the source in the figure for Prob. 1 1.8 using
a Y-A or A-Y transformation to simplify the circuit. From this result find the
Z_ Z
ab
a + Z^Z
Zj
z
Z Za
c
r
-)-
c
Z„Zj + z„z + z z a c c
Zbc
za
za z + z*z + z z„
Z ra ft c c
or in words:
by using one wattmeter for each of the three phases. This is illustrated for the three-
wire Y-connected load in Fig. 13. 8. Each wattmeter has its current coil in series with
1
one phase of the load and its potential coil across one phase of the load. The connec-
tions are theoretically correct but may be useless in practice because the neutral point
\ may not be accessible (as, for example, in the It would
case of a A-connected load).
be better, in general, to be able to make
measurements using only the lines a. h, and
the
c. In this section we shall show that this is, in fact, possible and that, moreover, only
two wattmeters are required, instead of three. The method is general and is applicable
to unbalanced as well as balanced systems.
n
Z|
/ \
b o z2
/ \
c o Z 3
f 1
'
Let us consider the three-wire Y-connected load of Fig. 13.19, which has three
wattmeters connected so that each has its current coil in one line and its potential coil
between that and a common point x. If T is the period of the source voltages and
line
i
B ,and ic are the time-domain line currents, directed into the loads, then the total
/',,,
a o-
rrcr\±
b o- < N
c o-
V ax — V aN I
V Nx
V bx = V bN + V Nx
v cx = v cN + V Nx
Thus the sum of the three wattmeter readings is precisely the total average power
delivered to the three-phase load, since the three terms in the integrand of (13.27) are
the instantaneous phase powers.
Chap. 13 Three-Phase Circuits 355
Since the point v in Fig. 13.19 is arbitrary, we may place it on one of the lines. Then
the meter whose current coil is in that line will read zero because the voltage across its
potential coil is zero. Therefore the total power delivered to the load is measured by
the other two meters, and the meter reading zero is unnecessary. For example, the
point x is placed on line b
in Fig. 13.20. and the total power delivered to the load is
P = PA+ Pc
where P A and Pc are the readings of meters A and C. It is important to note that one or
the other o\~ the two wattmeters may indicate a negative reading, and thus the sum
A
a o Z|
1 k °
+
A
! 1
B
bo
Do z2
i I
JUUU+
C "
+ nnnr
) '
o Z3
The proof of the two-wattmeter method of measuring the total three-phase power
has been carried out for a Y-connected load. However, it holds also for a A-connected
load, as the reader is asked to show in Prob. 13. 19.
As an example, in Fig. 13.20 let the line voltages be a balanced abc sequence with
z, =z =z = 2 3
io -fyion
Then we have
-V 6c
= -10(^/57- 120°:= 100-yT/60 o V rms
:
(100^J/^y)/-30 3
\ cC = 5^/T7-315° A rms
356 Three-Phase Circuits Chap. 13
= 317W
and
Pc = V I c6 1|
I cC |
cos (ang \ cb - ang I cC )
= 1183 W
with a total of
P= 1500 W
As a check, the power delivered to phase A is
= 100^/T
( >j )(
5 vr2")cos(-30° + 75°)
= 500 W
Since the system is balanced, the total power is
p=3Pp = 1500 W
which agrees with the previous result.
EXERCISES
13.5.1 In Fig. 13.18, let Z, = Z 2 = Z = 10/30^ fi, and let the line voltages be a
3
balanced abc sequence set, with \ ab = 200 /CT V rms. Find the reading of each
13.5.2 If the power delivered to the load of Ex. 13.5.1 is measured by the two watt-
meters A and C connected as shown in Fig. 13.20, find the wattmeter readings.
Check for consistency with the answer of Ex. 13.5.1.
Ans. 2/^/T, 4/^3" kW
13.5.3 Find the wattmeter readings PA and Pc and the total power P if the line volt-
ages are as given in Ex. 13.5.1 and Zj =Z =Z =
2 3
10/75° Q. Check the
answer by using P = 3PP . Ans. -598, 1633, 1035 W
Chap. 13 Three-Phase Circuits 357
PROBLEMS
13.1 It ym = \ nb = 100/0_° V rms, Z, = 10/30° Q, in /f-N. and Z 2 = 20/60° Q,
in A'-/?, find the neutral current \ nN in Fig. 13.4.
13.2 In Fig. 13.5, let V, = 100 /(T V rms, Z, = 5 + ;5 ft, Z 2 = 0.5 Q, Z 3 = 1 ft,
and Z4 = 10 — y"5 Q. Find the average power absorbed by the loads, lost in the
lines, and delivered by the sources.
13.3 In Fig. 13.10, the line currents form a balanced, positive sequence set with
\ aA = 10/0° A rms. If Zp = ^J [W Q, find the power delivered to the three-
phase load and the line voltages.
13.5 Repeat Prob. 13.4 if the load is A-connected and the phase impedance is
12/60° Q.
13.6 A balanced Y-Y system has a positive sequence source with \an = 100/0_° V
rms and Zp = 10 /30° Q. Find the line voltage, the line current, and the power
delivered to the load.
13.7 A balanced Y-connected source, \an = 240 /(T V rms, positive sequence, is
connected by four perfect conductors (having zero impedance) to an unbal-
anced load, Z AN = 10 -y'5fi, Z BN = 10 Q, and Z CN =j20Q. Find the four
line currents.
13.8 A balanced Y-Y three-wire, positive sequence system has Van = 200/0^ V
rms and Zp = 3 - y'4 O,
and the lines each have an impedance of 1 ft. Find
the line current and the power delivered to the load.
60 Hz and the line voltages are a balanced 200-V rms set, find the capacitances
required.
13.10 In Fig. 13.10, the source is balanced, with positive phase sequence, and
\ an = 100 /(F V rms. If the source provides 3 kVA at pf = 0.8 lagging, find
Zp .
13.13 A balanced Y-A system with V an = 120/CF V rms, positive phase sequence,
has Zp = 6 — j9Q. and an impedance of 1 Q in the lines. Find the power
delivered to the load.
13.14 A balanced three-phase, positive-sequence source with \ab = 200 [0^ V rms
is supplying a A-connected load, ZAB = 50 Q, Z BC = 20 -f-/'20fi, and
Z CA = 30 — y'40 Q. Find the phasor line currents. (Assume perfectly conducting
lines.)
358 Three-Phase Circuits Chap. 13
p(t) = v a ia +v b ib + vjc
is a constant, given by
13.19 Show that Px given by (13.26) is equal to the total average power delivered to
a A-connected load.
13.20 In the system of Fig. 13.20, the line voltages are a balanced, positive-sequence
set, with Va6 = 100/(r V rms, and Z, = Z 2 = Z = 3 10/3CT Q. Find the power
delivered to the load (a) by finding the readings of the two wattmeters and (b)
by P = 2PP .
13.22 Show that in Fig. 13.20 if Va6 = VJjt, VAc = VJa - 120° ,V ca = VL /a - 240°
Vrms, and Z = Z 2 = Z 3 = |Z| /60° then wattmeter A reads zero and watt-
x ,
T/2
P=
2|Z|
then the two wattmeters, connected as in Fig. 13.20, read, respectively, PA and
Pc where
,
P* _ cos (30° + 9)
Pc cos (30° - 9)
tan 9 = +/T
Pc+Pa
Thus we may find the power factor, cos 9, of the load from the two wattmeter
readings.
13.24 InProb. 13.23, find the pf of the load if (a) PA =Pc ,(b)P.4 = -Pc ,(c)PA = 0,
(d)Pc = 0, and (e) PA = 2PC .
14
COMPLEX FREQUENCY
AND NETWORK FUNCTIONS
In the previous chapters we have considered resistive circuit analysis, natural and
forced responses of circuits containing storage elements, and, in particular, ac steady-
state analysis.The excitations we have considered were, for the most part, constants,
exponentials, and sinusoids. In this chapter we shall consider an excitation, the damped
sinusoid, which includes all these excitations as special cases. From this function we
shall develop generalized phasors and general network functions which include, as
special cases, the phasors and impedances of Chapters 10-13.
The network functions will be expressed in terms of a complex frequency that
includes the frequency jco as a special case. The concepts of complex frequency and
general network functions will enable us to combine all of our earlier results into one
common procedure. Both the natural and forced responses of a circuit may be found
from its excitation and its network function, as we shall see. In addition, the network
function may be used to determine the frequency-domain properties of the circuit,
360
Chap. 14 Complex Frequency and Network Functions 361
v= Vm cos(cot I 0) (14.2)
v= V e" (14.3)
v=V (14.4)
2^-
dt
+ 5/ = 25e~' cos It (14.5)
+
3A AB = 25
-AA + 35 =
A/W
25 e"' cos 2t
V © 2H
or, cquivalently,
As expected, the forced response is. like the excitation, a damped sinusoid.
EXERCISES
14.1.1 Excite the circuit of Fig. 14.2 by the complex function
= 25e (
- 1+ J' 2 "
/, =(5 /-53.1° )g (
- 1+ ^"
~di\ ..
2^+5,, =«,
then Re ij is the response to Re v\. Apply this result to the functions of Ex.
14.1.1 to obtain /'given by (14.6).
14.1.3 Find the forced response v using the methods of Exs. 14.1.1 and 14.1.2 if
d3v ,d v 2
dv
+ 6 dF + U dt +6v =
, , ,
,
, ,
l
diT
where
i = 4e~ 2 cos<
(t - 60°)
Ans. 2e~ 2 cos'
(t + 30")
v x
= Vm cos (cot + 0) (14.7)
v x
= Rc(Vm e J
^e j0")
364 Complex Frequency and Network Functions Chap. 14
V= V„e'* = Vm [$ (14.8)
we may also write
A good question at this point is, Can we define phasor representations of damped
sinusoids that will work as well for us as the phasors, such as (14.8), of undamped
sinusoids? It seems plausible that we can, because the properties of sinusoids that
made the phasors possible are shared by damped sinusoids. That is, the sum or differ-
ence of two damped sinusoids is a damped sinusoid, and the derivative, indefinite
integral, or constant multiple of a damped sinusoid is a damped sinusoid. In all these
operations only Vm and <f>
may change. This is exactly the case with undamped sinu-
soids.
Let us see what follows if we write (14.10) in the form analogous to (14.9). That is,
identical in form to the damped sinusoid of (14.12). The only difference is that the
numbery'co is used in one case and the number s is used in the other. Obviously, then,
we may do anything with the damped sinusoids that we did with undamped sinusoids
as long as we use s instead of jco. We may define the phasor V of (14. 8) to be the phasor
representation of v in (14.10) and use it for damped sinusoidal circuit problems in
exactly the same way that we used it for the sinusoidal problems. In the damped
sinusoidal case we may wish to write the phasor as \(s) to distinguish it from the
undamped sinusoidal case, \(jco).
As an example, the damped sinusoid
where s = I
j2. Conversely, if V is given by (14.14) and s is as specified, then v
is given by (14.13).
Some authors prefer to call V(.v), corresponding to v(t) in (14.10), a generalized
phasor. even though it is identical to the phasors of sinusoidal functions. It is, how-
ever, a function of a generalized frequency, namely s, given by (14.11). Since s is a
complex number, it is more often called a complex frequency. Its components are
a = Re s Np/s
co = Im .s- rad/s
where the AT,, and s, are independent of /, may be said to be characterized by the
complex frequencies s u s 2 s n For example, writing (14.10) in the form
. . . . , .
v=Vm ^—-ti )
results in
v =K l
e
(a+Ja' u
+K 2e
{
°- Ja> ) '
sesses not one but two complex frequencies, namely Si = a +jco and s 2 = s* = a —jeo.
This concept of complex frequency is consistent with that of the natural frequencies
considered in Chapter 9.
EXERCISES
14.2.1 Find the complex frequencies associated with (a) 6 + 4e 2 ', (b) cos cot,
(c) sin (cot + 6), (d) 6e~ 3 sin (At + 10°), and (e) e~'(\ + cos It).
'
v=L*+Xt
is also a damped sinusoid of the same complex frequency.
L
Arts, v = Im s/(R + oL) 2
+ co 2 L 2 e°' cos (cot + + tan"" 1 °?
<f>
gL )
14.2.3 Find j and V(j) if w(f) is given by (a) 6, (b) 6e"', (c) 6e~' cos (2/ + 10°), and
(d) 6 cos (It + 10°).
/Ira. (a) 0, 6/CT; (b) -1, 6/0^; (c) -1 +j2, 6/10_°; (d);2, 6 /10°
14.2.4 Find v(t) if (a) V= 4/0°, 5 = -3; (b) V= 5/15°, ,y=/"2; and
(c) V= 6 / -30° s ,
= -2 +/4.
/Ira. (a) 4e~ 3 ', (b) 5 cos (2/ + 15°), (c) 6e~ 2t cos (At - 30°)
Because of the identical form of the phasor representations for sinusoids and damped
sinusoids, we may find forced responses to damped sinusoidal excitations using
phasors precisely as we did in the previous chapters. All the concepts and rules, such
as impedance, admittance, KCL, KVL, Thevenin's theorem, Norton's theorem, super-
position, etc., carry over to the damped sinusoidal case exactly. We need only to use
s = a -\-jco rather than/co.
It follows that, in the s domain, the phasor current l(s) and voltage \(s), associated
with a two-terminal device, are related by
\(s) = Z(s)l(s)
where Z(s) is the generalized impedance, or simply impedance, of the device. We may
obtain Z(s) from Z(jco), the impedance in the ac steady-state case, by simply replacing
jco by s.
Z*(*) =R
In the case of an inductance L, the impedance is
Z L (s) = sL
and for a capacitance C, it is
Chap. 14 Complex Frequency and Network Functions 367
Y*(*)
R
G, Yl(s) = IL>
Yf(, ° y<
The impedance seen from the source terminals consists of the impedances 2s and
5 in series and thus is given by
Z(s) = 2s + 5 ft
I(s)
5
Vg (S) f-\ 2s
Vg (s) = 25((F\
we have
V,(J) 25/0°
Ks) 5/-53.TA
2s + 5 2(-l+j2) + 5
As another example, let us consider the time-domain circuit of Fig. 14.4(a), where
it is required to find the forced response v (t) for a given damped sinusoidal input
v g (t). The phasor circuit is shown in Fig. 14.4(b), from which we have the nodal
equations,
(y + 1 + |*)V, - |v - V - ±.sY
g 2 =
+ -H v - v, = o
>
We note that V 2 = V /2 since the op amp and its two connected resistors constitute
a VCVS with a gain of 2. Therefore eliminating V, and V we have 2 ,
f
2A
v o—VW —VW— *
<i o "0
(o)
v
—w—^vw^
,, oV„
(b)
16
V (s) = V.(J) (14.15)
s
2
+ 2s + 8
'
If we have
v g (t) = e~ 2 cos At
'
V
\ (s) = s/T/\35°V
and therefore
Vo (t) = J~2e~ 11 cos (4/ + 135°) V
Chap. 14 Complex Frequency and Network Functions 369
As a final example in this section let us find the forced response i in Fig. 14.5 if
v gi = 8e"' cos / V
and
i
f2
= 2e-"A
/ = /,+/,
where i, is due to vel acting alone and i 2 is due to ig2 acting alone.
The phasor circuits for i g2 = and v gl — are shown in Figs. 14.6(a) and (b),
(o) (b)
respectively. The phasor currents are I, and I2, as shown. In Fig. 14.6(a), using current
division, we have
8/0° 4/s
I,
1 - {[(2s - 4)(4/j)]/(2* - 4 + 4/s)} 2s -r 4 + 4/s
I, - 2 va~/-45°A
Therefore the forced component due to v gl alone is
z !
= 2 /VA2V' cos (t - 45°) A
370 Complex Frequency and Network Functions Chap. 14
t
2
_ / [K4/J)]/(1 + 4/s) \
(2/0
o.
I2 = 0.8/0^ A
Therefore the forced component due to i g2 alone is
i2 = 0.8e" 5 'A
i = 2 v r 2V' cos {( -
/ 45°) + 0.8e" 2 '
A
EXERCISES
14.3.1 Change the voltage source in Fig. 14.5 to
v gl = 6e- 3 'cos2?V
and find the forced response /'t due to v gl alone. Arts. —4e~ 3 '
cos 2t A
14.3.2 Replace all the circuit of Fig. 14.6 (a), except the inductor, by its Thevenin
equivalent, and find I,.
32
Voc = ^74, „ 4(s + 5)'
= -1 +yl
th =
\r v
a
Arts. Z ,
,
where s 1 , -i
^ 4
14.3.3 Assume Ij = A 1 in Fig. 14.6 (a), and use the method of proportionality to
find the true I,.
^'- ''
=
,»+%'+ 10
= 2 ^=4g A
is the network function if I(s) is the excitation and V(s) is the response. On the other
Chap. 14 Complex Frequency and Network Functions 371
hand, if V(.v) is the input and \{s) is the output, then the input admittance
\( S ) - M
is t lie network function.
The foregoing examples are special cases m thai both the input and output are
measured at the same pair of terminals. In general, for a given input current or voltage
at a specified pair of terminals, the output may be a current or voltage at any place in
the circuit. Thus the network function, which we designate in general as H(.v), may be
a ratio of a voltage to a current (in which case its units are ohms), a current to a voltage
(with units of mhos), a voltage to a voltage, or a current to a current. In the last two
cases, H(.v) is a dimensionless quantity.
As an example, consider the circuit of Fig. 14.4(b), which was analyzed in the
previous section. If V„(i) is the output phasor and V g (s) is the input phasor, then, by
(14.15), the network function is
\ g (s) s
2
+ 2s H 8
As another example, if the source \g = 8/0° V is the input and I, is the output of
Fig. 14.6(a). then from Ex. 14.3.3 we see that
HM = = (I4I6)
ffi) ,' + fe + io
In the former example H(s) is dimensionless and in the latter case its units are mhos.
The network function H(.s) is independent of the input, being a function only of
the network elements and their interconnections. Of course, when the input is specified
this in a given application. From a knowledge of
determines the value of s to be used
the network function and the input function, we may then find the output phasor and
subsequently the time-domain output. To be specific, suppose V,(^) is the input and
\ {s) is the output. Then
from which
V.(5) = Ws)Yfc) (14.18)
where H(s) | |
is the amplitude and 6 the phase of H(s). Therefore if
Thus the amplitude of V is Vm H(s) and its phase is <f> 9. That is, the amplitude
\ |, +
of the output is that of the input times that of the network function, and the phase
of the output is that of the input plus that of the network function.
As an example, I,(.y) in (14.16) is given by
I,(j) = H(j)V,(j)
where V g (s) = 8/0° V and s = —1 + y I, as given in the previous section. Therefore
H(-l +jl) =
(-1 +yi) 2 + 6(-i +j\)+ 10
T7-45
and
= 2 <N /27-45° A
as was obtained earlier.
As a final note in this section, if there are two or more inputs, we may use super-
position to define a network function relating the output to each individual input,
with the other inputs made zero. For example, in the circuit of Fig. 14.5, we may find,
from Fig. 14.6(a) (with I 2
= 0), the network function
H,W - y
si
where s = -1 +./1 and V„ = 8/TV. Then from Fig. 14.6(b) (with V„ - 0), we
have the network function
H (j) =
2 ^
where s = —5 and \ gl = 2/0° A. We may then find Ij and I 2 and subsequently the
corresponding time-domain functions, /, and /,. These are, by superposition, the com-
ponents of i, as before.
EXERCISES
14.4.1 Given the network function
2( + 4)
HW- s2 +
' V
2^ + 2
and the input
y.(s) = 4/0°
—
Chap. 14 Complex Frequency and Network Functions 373
find the forced response r (t) if (a) s = — 1 +73, (b) s = —3 + j\, and (c)
s= -1. Ans. (a) 3VT <r' cos (3? - 135°), (b) -2e" 3 'sin /, (c) 24<r'
14.4.2 Find H(.v) if the response is (a) !,(.?), (b) \ (s), and (c) \ (s).
Ans. (a)
J2 + 3s + 1
(b)
1 1
vw
(c)
(s + l){s + 2)(s I
;,
3)'
v "'
(^ | l)(s + 2)(s + c,
3)' (^ + l)(j + 2)
3A IH
—vw
i
tw^
+
v
»©
EXERCISE 14.4.2
In general, the network function is a ratio of polynomials in s with real coefficients that
are independent of the excitation. To illustrate this, let us consider the example of
Fig. 14.2, described by
2^ + 5/ =v
dt
where i is the output and v is the input. Using the same technique with complex forcing
we used Chapter we note if v = We then the output must 1
functions that in 10, that ',
have the same form, namely i — le", where, of course, V = \(s) and I = l(s) are the
phasor representations of v and /'.
Substituting these values into the differential equa-
tion, we have
(25 + 5)Ie" = \e"
V 2s
In the general case, if the input and output of the circuit are v,(t) and v (t), respec-
tively, then the describing equation is
'
a.
d"v
cit"
a .-1
dn
dt
n-l
]
v
a^ + a v Q
dm v l i\
+ *.-. T *=r + l
i6,f + V, (14.22)
The tf's and fe's, of course, are real constants and are independent off
374 Complex Frequency and Network Functions Chap. 14
As before, if v t
= \e t
s
', then the output must have the form v =V e", where
V,(^) and V (5) are the phasor representations of v t
and v . Substituting these values
into (14.22), we have
(a n s
n
+ flB -iJ
B_1
+ • • •
+ OiS + a )\ o e"
= {b m s
m
+ 6m . 1
5- + 1
• • •
+ b s
x
+ & )V,e" (14.23)
Vote)
VXj)
_ HM " sm + 6
w - b«„*" + a„. "
"--'-y
1
'""
5"- 1
1
+
+ •
•
•
'
'
+
+ ^5 +
a lS +
bQ
a
Kl ^
(W)A\
V
which is a ratio of polynomials in s.
We may also write the network function (14.24) in the factored form
hw = ^"
- Pi)(s —p
i?:"?!
,,,
— p„)
!rn"l
tf n (s
zi
2) . . . (s
(i4 25)
-
In this case thenumbers z 1 ,z2 ,...,zm are called zeros of the network function because
they are values of s for which the function becomes zero. The numbers p lt p 2 ,p„ ,
are values of ^ for which the function becomes infinite and are called poles of the
network function. The values of the poles and zeros, along with the values of the
factors a„ and b m uniquely determine the network function.
,
m K
v
}
_ 6(5
s(s
+ l)(s + 2^ + 2)
+ 2)(s + 4^+13)
2
2
As we shall see in Chapter 15, pole-zero plots are useful in considering frequency-
domain properties of circuits.
Chap. 14 Complex Frequency and Network Functions 375
jOO
2+ j3 x -3
2
-i + i"o I
o
-2 -I
-l-jlO -I
-2
-2-j3 x -3
EXERCISES
14.5.1 If the zeros of H(s) are s = —1, —1 ±/'l, the poles are s = —2, —1 ±j2,
and H(0) = 4, find H(s).
Ans
20(s + 1)(^ + 2s + 2)
-
{s + 2)(j 2 + 25 + 5)
14.5.2 Draw the pole-zero plot of the network function of Ex. 14.5.1.
As we know from our previous experience, an output of a circuit consists of the sum
of a natural and a forced response. In the last few chapters we have been concerned
entirely with finding the forced response, and we have seen that the phasor technique
enables us to do this in a very easy, straightforward way in the cases of sinusoidal,
damped sinusoidal, and exponential excitations. In power systems studies the forced
response is, of course, an ac steady-state response and is always present. Therefore the
forced response is usually of more interest than the natural response, which is transient
and gone after a very short time.
With damped sinusoidal excitations, on the other hand, both the forced and
natural responses are transients. (In an actual circuit the natural response must be
a transient, for otherwise we would have either a sustained or a growing response
without an external excitation.) Therefore the natural response assumes more impor-
tance, relative to the forced response, than it does in ac steady-state cases. In finding
376 Complex Frequency and Network Functions Chap. 14
K } ~ V,(.y)
~ D(s)
(14.27)
N(s) = ^ b m s m + b m ^s m
~ l
+ • • •
+b s + x
b (14.28)
The last two expressions are merely the right member and left member, respectively,
of (14.22) with the derivatives replaced by powers of s. For example, if the describing
equation is
fe + A d-^- +
1
3v = 2* +v t
(14.30)
dt dt dt
H(KS}
S) = Y£?} = 2s + l
(14
K 31)}
V,(s) s
2
+ 4s + 3 •
D(s) = (14.32)
1
d"v„ ,
d"~ v . . dv, .
„, a
roots are the natural frequencies of the circuit. Since these roots, by (14.29), are also
Chap. 14 Complex Frequency and Network Functions 377
the poles of the network function, we see that the natural response of the circuit is
vn = A x
e** + A 2 ep " + • • •
+ A n e" (14.33)
where the natural frequencies/),. />, pn are the poles of the network function and
A lt A 2 4„ are arbitrary constants. Modifications, of course, must be made, as
described in Chapter 9, if the natural frequencies are not distinct.
We now have a very simple method, based on phasors, for finding the complete
response of a circuit. All we need to find is the network function from which by (14.27)
we may obtain the output phasor. The forced response is found from the phasor
Therefore since V,(.v) has no poles, the poles of H(s) are the poles of V (^). Thus the
complete response v (t) may be obtained from its phasor representation \ (s). The
forced response is obtained as before, and the natural frequencies are the poles of
y„(s), from which the natural response may be constructed.
As an example, let us find the complete response /'
of the circuit of Fig. 14.8. Using
current division, the phasor I(s) of i is given by
=
\ e {s) 3s
!(*)
12 + {{3s{2s + 6)]/(3j + 2s + 6)} 3s + 2s + 6
(s + \)(s+ 12)
since Vg = 2/0°. Since the poles of I are s = — 1, — 12, the natural response is
i2n
2t
2e" cos t ^4>\ 3H 2H
1 = 0.16 /12.7°
so that the forced response is
/(,) =A x
e-' + A 2 e~ xlt + 0.16e" 2 cos '
(t + 12.7°) A
The arbitrary constants may be evaluated, as in Chapter 9, if we know the initial
energy conditions.
EXERCISES
14.6.1 Find the natural response in Ex. 14.4.1, assuming that there is no cancellation
in the network function. Ans. e~\A cosx
t + A2 sin t) V
14.6.2 Find the complete response in Ex. 14.4.1 (a) if the natural response is as given
in Ex. 14.6.1 and
Wo(0+)==
«n =
Ans. 3e"'[cos / - 3 sin + «/2 cos (3r - 135°)] V
/
14.6.3 Find the complete response v (t) in Ex. 14.4.2 if v {t) = e~' cos V and t
g
lo{K)
'
~ dt ~ dt 2 ~ u
(Note that the natural frequency s = —3 has been cancelled in the network
function. Thus the corresponding term must be added to the natural response.)
Ans. -le~' + le'
1' - \e~*' + (1/V2>~ r
cos (/ - 135°) V
As we have seen, the natural frequencies are the poles of the network function if
For example, let us consider the circuit of Fig. 1 4.8 with the source killed (since
the natural response corresponds to a zero source). The natural frequencies are the
poles of any network function. Therefore we may some proper
excite the circuit in
manner and, some chosen output, determine the network function. Figure 14.
for
illustrates the two proper means of applying an excitation to the dead circuit. We may
insert a voltage source in series with an eicment as in x-x' of Fig. 14.9(a), or we may
place a current source across an element, as across y-y' of Fig. 14.9(b). The first entry
into the circuit is sometimes called the pliers entry since we cut a wire and insert a
voltage source. The second entry is a soldering entry since we solder the source across
two nodes.
(a)
In the pliers entry, if V^s) is the voltage source inserted and I x (s), the current into
the circuit at the source, is the response, then the network function is
Y,(J)
Us)
V x (s)
Thus the natural frequencies are the poles of \ x {s) or the zeros of its reciprocal, Z^),
the input impedance at x-x'. In this case we have
12(3j)
Z x (s) = 2s + 6 + 12 + 35
Z,(*) yx*)
Us)
Therefore the natural frequencies are the poles of Z y (s) or the zeros of its reciprocal,
Y y (s), given by
380 Complex Frequency and Network Functions Chap. 14
(s + \)(s + 12)
125(5 + 3)
Again, the natural frequencies are — 1, —12.
EXERCISES
14.7.1 Find the natural frequencies of the circuit of Fig. 14.5 by killing the sources
and using (a) a pliers entry in series with the capacitor and (b) a soldering entry
across the capacitor. Arts. —3 ±jl
14.7.2 Find the natural frequencies of the circuit of Ex. 14.4.2 by killing the source
and using (a) a pliers entry in series with the capacitor and (b) a soldering entry
across the capacitor. (Note: Do not cancel the common pole and zero,
s = —3.) Ans. —1, —2, —3
for which the input and output measured at different pairs of terminals.
signals are
The simplest and most-often-encountered circuit for which this is possible is the two-
port network, a port being defined as a pair of terminals at which a signal may enter
or leave. A general two-port network is symbolized by Fig. 14.10(b), as contrasted
with the one-port network of Fig. 14.10(a).
In general, as seen in Fig. 14.10(b), a two-port network has four terminals. It is
possible, of course, for two of the terminals to be the same, in which case we have a
three-terminal network. A general example of this case is shown in Fig. 14. 1. 1
We may associate two pairs of currents and voltages with a general two-port
network, as shown in the frequency-domain case of Fig. 14.12, with variables V^s),
I,(s), \ 2 {s), and l 2 (s), as indicated. In case the network is linear, these variables may
be related in a number of ways. For example, if l and I 2 are inputs and V! and V 2
l
are outputs, we know by superposition that V, and V 2 each have components pro-
o o o
o o
o o o
(o) (b) o * o
FIGURE14.10 (a) One-port and (b) two-port network FIGURE 14.11 Three-terminal
two -port network
;
V, z 1
1
' i i
z i 2 '
(14.34)
V, *2|I| z 22':
where the z's are proportionality factors, which in general are functions of s.
Since the z's multiply currents to yield voltages, they must be measured
in ohms.
_V,
I, I2 =
(14.35)
i,
_V,
I, =
(14.36)
II I
2
z Z2
+ l
v l
z3
— i
i
— Z,
*2I Z3 (14.37)
z,
If in Fig. 14.12 Vj and V 2 are inputs and I, and I2 are outputs, we may obtain, in
an analogous manner,
1
(14.38)
*2 == y21* 1 ~T~ V 22'2
where, evidently,
y.r
y !2
v 2
I2
y 2 i-y^
I2
y22 — tt
v, = o
yiI =Y +Y a 4
yi2 =y 2 (14.39)
Y +Y 6 c
The z- and >'-parameters are but two sets of parameters associated with a two-port
network. Others include the hybrid parameters and the transmission parameters, which
willbe defined and considered in the exercises and problems.
The two-port parameters may be used to obtain various network functions. For
example, if port 2 is open (I 2 = 0), then from (14.34) we may find the voltage ratio
function
Xa = hi (14.40)
v 1
z i 1
As another example, let us find the voltage ratio function for the two-port loaded
with 1 f2, as shown in Fig. 14.15. Evidently, we have V = —
2 2, which substituted into
T 'b T °
Ya Yc
o • * o
v 2= -y 2
V, " + 1 y2
,
(14.41)
Another use for the two-port parameters is in the construction of equivalent cir-
cuits. For example, in the first equation of 14.34), V, is a sum of two terms, z, ,1, and
(
z !2 I : The first may be obtained by an impedance z M carrying a current I,, and the
.
may interpret the second equation of (14.34) and draw the equivalent circuit. The
result, as may be verified by inspection, is shown in Fig. 14.16. We say that this circuit
is equivalent (that is, at the terminals) to that of Fig. 14.12, because they both have
the same two-port parameters. Thus if the port currents are the same, then the port
voltages are the same, and vice versa, for both networks.
If z 12 = z 21 (or, equivalently, y
l2
= y 21 ), the network is said to be a reciprocal
network. If the elements comprising the network are resistors, inductors, capacitors,
or independent sources, then the network is always reciprocal. An equivalent circuit
for a general reciprocal three-terminal network shown in Fig.
is 14.17. There are, of
course, many other equivalent circuits that may be drawn using the various network
parameters.
Z II" Z I2 Z 22~ Z I2
+
Z|2 v2
o < I
EXERCISES
14.8.1 Draw an equivalent circuit for a two-port network with parameters z u = 6 Q,
z i2 = z 2i = 4 Q, and z 22 = 10 Q.
14.8.2 Show that the ^-parameters may be obtained from the z-parameters by
Z^ > Zi i
Zi i Zi i
where A = z n z 22 — Zi2 z 2i- Use this result to find the ^-parameters of the
circuit of Ex. 14.8.1. [Suggestion: Solve (14.34) for the voltages, and compare
the result with (14.38).] Ans. y n = ^, y 12 = y 2 i
= - 77, Y22 = ^U
.8.3 Two sets of hybrid parameters, h n , h 12 , h 21
2I , h 222 and g u g 12) g 21 g 22
, , , are
defined by
V, =^,1, +h 12 V 2
I2 = h 2 ili + h 22 V 2
and
Ii =guVi + g ia I a
V 2 =gaiV, + g 22 I2
Find the A-parameters and the ^-parameters of the network of Ex. 14.8.1.
„-„. 1 _4 4 44
PROBLEMS
14.1 Find the phasor representation \(s) of (a) v(t) = be'' cos {At + 30°), (b)
14.2 Find v(t) if (a) \(s) = 6/10° with j = -3 + j 8, (b) VQy) = 5/0° with s = -10,
and (c) V(s) = 4 + /3 with 5 = -1 1/2.
14.3 Find the complex frequencies which characterize (a) v = e~'(cos It + sin 2/)
and (b) v = e~ l <
+ e~ 4 ' + 2 cos t.
14.4 Find the impedance Z{s) and locate its poles and zeros. If an excitation
vg = 6e~' cos It V is applied at the terminals, find the forced component of
the current it delivers.
Chap. 14 Complex Frequency and Network Functions 385
in I H
AMA-
Z(s) F -±- i a
PROBLEM 14.4
60
—W^~
Z(s)- -4 F ~7 F = Fi
F
PROBLEM 14.5
his) U(s)
H,
V,(j)
H 2 (.v)
~ v '
'
V ? U)
14.8 If the voltage source in Prob. 9.18 is r e (t) = 6e~' cos It V, find the network
function l(s)/V g (s) and the forced response i in each case.
14.9 In Prob. 14.7, replace everything in the circuit except the 5-H inductor by its
Thevenin equivalent circuit. Use the result to find the network function and
the forced response.
14.10 In Prob. 9.19, let the current source be i g (t) = Ae' 1 cos At
'
A and C= \ F.
Replace the corresponding phasor circuit, except for the series combination of
1 H and 1 Q, by its equivalent Thevenin circuit. Use the result to obtain the
network function I(s)/l g (s) and the forced response /(/).
14.11 Find the network function \(s)/\ g (s) in Prob. 9.21. Use the result to find the
forced response dt).
14.12 Find the network function \(s)IV g (s) in Prob. 9.22. Use this result to find the
forced response v(t) if v g (t) = 6e~ 2 ' cos At V.
14.13 Find the network function and the forced response v(t) if v g U) = 6e~ 2 V.
'
386 Complex Frequency and Network Functions Chap. 14
in h h
—VW—
i
i
Offiff
v-T rffWv
i
PROBLEM 14.13
H(*)
V (s) 4s(s + 2)
V,(s) (* + !)(* + 3)
If no cancellation has occurred, find the complete response v (t), for / > 0, if
v t (t) = 6e~' cos It V and v (0 + ) = dv o (0 + )/dt = 0.
14.15 Repeat Prob. 14.14 if
H(s)
4s(s + 2)
s2 + 2s + 2
14.16 Repeat Prob. 14.14 if i>,-(f) = 6e ' V. (Suggestion: Observe the describing differ-
ential equation.)
14.17 Find the complete response ?>(/), for t > 0, if there is no initial stored energy.
PROBLEM 14.17
14.18 Find the network function V (j)/V,-(*) and the forced response if (a) v t (t)
= 6e~ 2t cos / V and (b) v (t) = 6 cos / V. t
14.20 Find the network function, the natural frequencies, and the forced response
v if i
g
= 5e~ 2 cos 2/ A.
'
14.21 Find the network function \(s)/\ g (s) in Prob. 9.23 and the forced response
v(t) if v g (t) = 6e~ 2 cos
' / V.
14.22 Find the complete response v(t), for / > 0, in Prob. 14.21 if there is no initial
stored energy.
14.23 Considering the figure for Prob. 14.17 as a two-port network with terminals as
shown, find the z-parameters as functions of s.
Chap. 14 Complex Frequency and Network Functions 387
PROBLEM 14.18
PROBLEM 14.19
PROBLEM 14.20
14.24 Show that the hybrid parameters defined in Ex. 14.8.3 may be obtained from
the z-parameters by
A *12
h h 12 =
z2 2
h 21 = -m. h 22 = J_
*2 2 Z2 2
gi; gl2 =
o
62 1
-5*1,
—~Z ll
' 822 = —A
z ll
14.25 Find the y-, h-, and ^-parameters of the two-port network of Prob. 14.23.
.
V, = AV 2 - BI 2
I, = CV - 2 DI 2
Show that
B
Z21
C= —
Z:i
D Z 22
Z2I
where A is given in Prob. 14.24. Note that the transmission parameters relate
the output variables V2 , I 2 to the input variables V,, I,.
14.27 Find the transmission parameters of the two-port network of Prob. 14.23.
14.28 Derive the Y-A transformation of Chapter 13 by making Figs. 14.13 and 14.14
equivalent at the terminals (that is, equal two-port parameters).
14.29 Find the ^'-parameters of the network shown. Ferminate the output port with
a 1-fi resistor, and find the resulting network function V 2 /Vj.
i,/2 V
in
J VW-
I F
PROBLEM 14.29
14.30 Show that for Fig. 14.15, in terms of the transmission parameters we have
V, 1
V, A +B
14.31 Check the result for the terminated network of Prob. 14.29 by using the result
of Prob. 14.30.
15
FREQUENCY RESPONSE
Frequency-domain functions are very useful, as we have seen, for finding corre-
sponding time-domain functions. The frequency response of a circuit, however, is
extremely useful in its own right, as we shall see in this chapter. For example, if we are
interested in which frequencies are dominant in an output signal, say Y(jco), then we
need only consider the amplitude |V(yo))|. The dominant frequencies correspond to
relatively large amplitudes, and frequencies that are virtually suppressed correspond to
relatively small amplitudes.
There are many applications in which frequency responses are important. One
very common application is in the design of electric filters, which are networks that
pass signals of certain frequencies and block signals of other frequencies. That is, if the
output signal of the filter has amplitude |
V(y'co) |, then co, passes if |
V(y'«,) |
is relatively
large and is blocked if V(y"a>,) is relatively small. There are many examples of electric
| |
filters in our modern society, some of the more common being those in our television
sets, which allow us to tune in a certain channel by passing its band of frequencies
while filtering out those of the other channels.
In this chapter we shall consider frequency responses, both amplitude and phase.
We shall also define resonance and quality factor and show how they are related to the
frequency responses. Finally, we shall consider methods of scaling networks to yield a
given frequency response with practical circuit element values.
389
)
As we know, we may also write the network function in the polar form
where H(jco) | |
is the amplitude, or magnitude, response and 0(co) is the phase response,
given, respectively, by
HO'cj)| = VRe 2
H(j(o) + Im 2 H(j<o) (15.3)
and
= Im H(jco)
0(a)) tan i
(15.4)
Re H(jco)
The amplitude and phase responses are, of course, special cases of frequency responses.
As an example, suppose the network function of the RLC parallel circuit of Fig.
15.1 is the input impedance
H(s)
v 2 (.)
Z(n) _ l
Us) -
"*> sC
(15.5)
(\/R) I | (IM)
or
Ks)-
1
'?* '',_ (15.6)
, , "i
HO) =
(l/R)+j[coC-(l/coL)]
I
(15.7)
Hijco) (15.8)
V(l/fl 2 H [a)C-(l/a)L)] 2
and
Since R, L, and C are constants, the maximum amplitude occurs at the frequency
to = a) for which the denominator in ( 1 5.8) is a minimum. Evidently this occurs when
Chap. 15 Frequency Response 391
coC - -V =
or
H(ju>)
/lc
(q) (b)
then the input phasor is I = /m /0°, and the output phasor is V2 = Im Z = / m H. Thus
the amplitude of the output is simply that of the network function multiplied by a
constant. Therefore we may obtain as much information from the network function
response as from the output response. For this reason and the reason that the network
function depends only on the network and not on how it is excited, we shall usually
consider the frequency response of the network function.
EXERCISES
15.1.1 Let R = 2Q, L = \ H, and C = \ F in Fig. 15.1, and find the maximum
amplitude and the point where it occurs. Also, sketch the amplitude and phase
responses. Ans. |H| max = 2, co
392 Frequency Response Chap. 15
15.1.2 For the RLC series circuit with a voltage source v g , let the network function
be H = I/Vg, where I is the phasor current. Show that the amplitude and phase
responses are similar to those of Fig. 15.2 with |H| max = l//?andco = l/*/LC.
15.1.3 Let the network function of Fig. 15.1 be H= IJli, where I L is the inductor
phasor current directed downward. Show that
\\LC
H(.r)
s* +(\IRC)s + (1/IC)
2R 2 C<L
15.2 FILTERS
With reference to Fig. 15.2(a) we see that frequencies clustered around co = \/*/LC
rad/s, or/ = \/(2n^/LC)Y\z, correspond to relatively large amplitudes, while those
near zero and larger than co correspond to relatively small amplitudes. Thus Fig. 15.1
is an example of a bandpass filter, which passes the band of frequencies centered around
eo . In the general amplitude case, shown in Fig. 15.3, we say that co , the frequency at
'I
' 2
which the maximum amplitude occurs, is the center frequency. The band of frequencies
that passes, or the passband, is defined to be
co Cl <co < co t ,
where and co are called the cutoff points and are defined as the frequencies at
co Cl c ,
which the amplitude is l/^/T = 0.707 times the maximum amplitude. The width of
Chap. 15 Frequency Response 393
B= o) c <°c, (I5.ll)
2A
v.o V\AA
V 2 (j) _ 2
His)
V,(J) 2s + 2
H(;w)|
V(2 — co
2
)
2
+ 4cu
2
H(;co)j (15.12)
VI -(co* 14)
H(joo)
(a) (b)
1
H(M)I HOa>)|
(1)
VI +(co 4 /4) c
whose only real answer is co c 2. Thus the band of frequencies which passes is the
low-frequency band
< (o < ^fl
The phase response for Fig. 15.4 may be easily shown from H(jco) to be
2co
<f>(co) = — tan" z
co
EXERCISES
15.2.1 Show that
2v
H(.v)
s2 + 0.2s + 1
H(.v)
3(.v
2
+ 25)
s + 25
15.3 RESONANCE
A physical system which has a sinusoidal type of natural response reacts vigorously,
and sometimes violently, when it is excited at, or near, one of its natural frequencies.
This effect may have been noticed by the reader in Sec. 9.6, particularly in the case of
Ex. 9.6.3. The system in this respect is somewhat like all of us. When urged to do what
it naturally wants to do, it responds with enthusiasm.
phenomenon is known as resonance, and its side effects may be good or they
This
may be bad. As an example, a singer may break a crystal goblet with his voice alone by
properly producing a note at precisely the right frequency. Also, a bridge may be
destroyed if it is subjected to a periodic force with the same frequency as one of its
natural frequencies. This is why no thoughtful troop commander will march his men
in step across a bridge. On the other hand, without resonance there could be no
electric filters.
We shall define a sinusoidally excited network to be in resonance when the ampli-
tude of the network function attains a pronounced maximum or minimum value. The
396 Frequency Response Chap. 15
frequency at which this occurs is called the resonant frequency. As an example, the
RLC parallel circuit of Fig. 15.1 is in resonance when the frequency of the driving
function is co = \\J LC. This was shown in Sec. 5. where it was demonstrated that 1 1
Fig. 15.2(a) is typical, with its peak at the resonant frequency. The
relatively high
parallel RLC circuit is so important that the term parallel resonance is reserved for its
resonant condition.
The reader may recall encountering the term resonant frequency earlier in Sec. 9.8
in connection with the underdamped case of the parallel RLC circuit. The two fre-
where
(15.14)
2RC
and
From the pole-zero plot, shown in Fig. 15.7, we see that the resonant frequency &> ,
the radius of the dashed semicircle. If R is made larger, then a is made smaller and the
— jCO
jw d
j^o
' l\
\
/
/ V \
/ \
\ (* t
1
f
1
-a a
\
\
M-~:
FIGURE 15.7 Pole-zero plot of the parallel RLC circuit
Chap. 15 Frequency Response 397
resonant frequency is even closer to the natural frequency. In this case, as is clear in
peak is more pronounced. Of course, if R were infinite (open-circuited),
Fig. 15.2(a). the
then the resonant frequency would coincide with the natural frequency and the
amplitude would be infinite.
Before leaving the parallel RLC circuit, let us note that the network function
is actually the input impedance, as stated in (15.5). The resonant frequency
co = co = 1/+/LC is the frequency for which the input impedance is purely real, as
seen in (15.7). Indeed, many authors define the resonant frequency precisely this way in
the case of a two-terminal network. In the general case, maximum amplitude does not
always occur exactly at the frequency of real impedance, but usually there is very little
difference.
In the case of the RLC series circuit excited by a voltage source \g , if the phasor
current I is the output, then
HO"*,) = #M
V
=
g
(jco)
Y(Jto)
KJWJ
~
R + j[coL -
l
(1/©C)]
where Y is the input admittance seen at the source terminals. Evidently series resonance
occurs at co = 1/+/LC, yielding a maximum amplitude of \jR. As in the parallel
resonance case, the resonant frequency is also the frequency of real input impedance or
admittance. At resonance the effect of the storage elements exactly cancels, and the
source sees only the resistance.
EXERCISES
15.3.1 Show that the resonant frequency in the case of the function of Ex. 15.2.1
coincides with the frequency for which the function is real.
15.3.2 Find the resonant frequency for the parallel RLC circuit described by (a)
R= 2 kQ, L = 4 mH, and C = 0.1 //F; (b) co d = 5 rad/s and a = 12 Np/s;
and (c) a = 1 Np/s, J? = 4 Q, and L = 2 H.
Am. (a) 50,000, (b) 13, (c) 2 rad/s
15.3.3 In Ex. 15.3.2 (a), find the amplitude of the voltage across the combination if
where K, a > 0, and b > are real constants. (The function is called second-order
because of the quadratic denominator.) To see that the function is of the bandpass
type, let us consider its amplitude
A'col
HO) |
J(b -
I
(o
1
)
1
-I- a 2 co 2
\K\
V« 2
I Kb - 2
(o )/coY
HOeo) m .* = J —
a
l
b = a>l (15.17)
-jY I
or
\K\ \K\
J a1 + [(b - 2
co )/co c Y J2a
which evidently holds if
b — co
2
= ±o ,
Thus we have
co
2
± aco c - b = (15.18)
which, because of the double-sign possibility, has four solutions. Using the positive
sign, we have, by the quadratic formula.
J
Q. t
= ~° +
f+ 4b
(15.19)
We have discarded the negative sign on the radical because this gives a negative cutoff
frequency. Using the negative sign in (15.18) and again suppressing the negative root,
we have the other cutoff frequency,
- U + V + 4b
(15-20)
<o ct
f
Evidently, from (15.19) and (15.20) we have the bandwidth,
B = co c , - co cl = a (15.21)
Chap. 15 Frequency Response 399
Thus in view of (15.17) and (15.21) we may write the network function as
S
H (*) = o 2 ( ' 5-22)
s
2
z
+,
Bs + g>o
which is the general network function of a second-order bandpass filter having center
frequency co and bandwidth B. The amplitude response is shown, of course, in Fig.
15.3.
Another result worth noting from (15.17), (15.19), and (15.20) is
col = co Cl co c , (15.23)
which demonstrates that the center frequency co is the geometric mean ^/co Cl co c , of
the cutoff points.
A good measure of selectivity or sharpness of peak in a resonant circuit is the so-
called quality factor Q, which is defined as the ratio of the resonant frequency to the
bandwidth. That is.
2= ^ (15.24)
(The letter Q is symbol for reactive power, as the reader will recall. However,
also our
the two quantities will in the same context so there should be no confu-
never be used
sion.) Evidently, since B = co /Q, a low Q corresponds to a relatively large bandwidth,
and a high Q (sometimes arbitrarily taken as 5 or more) indicates a small bandwidth,
or a more selective circuit.
With this definition of Q we may write (15.22) in the form
so that we see at a glance the center frequency, the quality factor, and the bandwidth.
As an example, the filter described in Ex. 5.2. 1 1 is a bandpass filter with co = 1 rad/s,
B= 0.2 rad/s, and Q = 5. Another example is the parallel RLC circuit with the
network function given in (15.6). In that case, co Q = \J«JLC, B = l/RC, and Q=
co IB, which has the equivalent values
Q- co RC
= R^ (15.26)
=~ _R_
oi a L
In the examples we have considered, the two definitions are the same, as the reader is
asked to show in Exs. 1 5.4. and 5.4.2. However, in general, there is a slight difference.
1 1
».....=
^!+V^ + (i)
or
Evidently if Q is high, then we may neglect (\J2Q) Z in comparison with 1, and write
©cc^T^+Wo
2Q
or, approximately,
co Cl = co - B ,
(15.29)
S
Thus as £> increases, the amplitude response approaches arithmetic symmetry. That is,
EXERCISES
15.4.1 For the RLC parallel circuit in resonance, show that the total energy stored is
w(t) = yR 2
CIi cos 2 o> t + ^| sin 2 co t
= ^ C/ 2 2
15.4.2 Show thai for the circuit of Ex. 15.4.1 the energy dissipated per cycle is
nRIl
RJ;„ cos 2 av dt a»>,
and thus by definition (15.27) and the result of Ex. 15.4.1 that
Q cOqRC
15.4.3 For the RLC series circuit with excitation vg = Vm cos cot V, show that
co = 1/v/IC and = co L//?.
The pole-zero plot of a network function may often be used to readily sketch the
frequency responses. To seehow this may be done, let us write the network function in
the form
His)
K(s - *,)(j -z 2
)...(s ,)
(15.30)
(s - p,)(.v —p 2 ). . .(s --/>„)
where, as before, the z's and and poles. Each of the factors is a
/>'s are the zeros
complex number of the form s —
and may be represented in the .v-planeby a vector,
5,
drawn from ,v, to as shown .v. in Fig. 15.8. This is true since by vector addition the
written
H(jco) = \H(jco)\e^'" )
H(/©)| = KN N *
2 • •
N™ (15 31)
H(j)
s
2
+ 2s + 401
which has a zero at and poles at —1 ±./20. These are shown on the pole-zero plots
of Fig. 15.9. By (15.31) and (15.32) we have
H(yo»)'
I
MM 1 2
<f>(w) = a - (/?, + fi t )
+
whose components are identified in Fig. 5.9(a). First we note that 1 if co varies from
to +oo, then a = 90°. For w = 0we have, from Fig. 15.9(b),
|M|-
1111 ^°>_ -Q
v/40V401
<j> = 90° - (-tan" 20 1
+ tan" 1
20) = 90°
_j*(20)_
'"'
(1)V'601
= 90° (0 | tan 1
40)^0
This point is in the region where the amplitude changes the fastest, since Af , = 1 is a
Chap. 15 Frequency Response 403
J20
y*.
MA
-I
f
^
M2 /
2
h\
-J20
(b)
(c) (d)
If co = co h , a very high value, say. such as 10 6 then , all three vectors will be essen-
tially vertical, so that
N ^ M ^ M ^ co
t 2 h
a ^ ^ £ ^ 90
/?, 2
Sketching the functions, we have the forms of Fig. 1 5.2, as expected, with co «= 20
and |H| max ^ 2.
We may rough idea of the cutoff points from Fig. 15.9(d). In this figure
get a
M x
= J~1
/ is ', which
,J~1 times its value at the approximate peak, represented by
Fig. 5.9(c). Since
1
2 M
and N have changed by much smaller percentages, the amplitude
is then approximately \\^fl times its peak, so that co Cl 19. By a similar argument at ^
co = 21, we have co C2 «=21. The exact values, by (15.19) and (15.20), are
ffl CliC = 19.05, 21.05. In this example, by comparing the network function with
.
(15.25), we see that Q = 10.01. Thus we have a high Q so that the poles are very
near the j'co-axis. Thus cod is very near a> and the approximations we have made are
,
EXERCISES
15.5.1 Use the method of this section to sketch the amplitude and phase responses
for
S* + 4s + 2504
15.5.2 Find the exact values for the answers given in Ex. 15.5.1.
Ans. 50.04, 12.51, 4, 48.08, 52.08
The reader may have noticed that in many of our examples we have used network
elements such as 1 CI, 1 F, 2 H, etc., which are extremely nice numbers to have when we
are analyzing a network. For example, in Fig. 15.4 we had elements of 2 Q, \ F, and
\ F, and the network was a low-pass filter with a cutoff frequency of ^J~2 rad/s, or
0.23 Hz. However, such element values as these are not very practical, to say the least,
and there is very little demand for filters that pass frequencies between and 0.23 Hz.
(As an illustration, a 1-F capacitor, constructed of two parallel plates 1 cm apart with
air as the dielectric, would require a face area of 1.13 • 10 9
m 2
.)
an impedance given by
Z'(s) = sL' + R! + ±
The network has been impedance-scaled by an impedance scale factor k t
if the imped-
ance Z(s) of the scaled network is Z(.v) = k,Z'(s). In other words, we must have
Zis)=k,(sL' + R'+±^ 1
Z(s) = sL +R+ -^
L --
= k tL' U
R =- k,R' (15.34)
C =
C
'
k,
may be shown to hold in general. Also, if there are dependent sources, the scaling is
accomplished by multiplying gain constants having units of ohms by A,, dividing those
with units of mhos by k and leaving unchanged those that are dimensionless. The
t,
scaling multiplies by A,., divides by A,, or leaves unchanged network functions that have
units of ohms or mhos or are dimensionless.
To illustrate, let us = 5. The 2-fi
impedance-scale the network of Fig. 15.4 by k t
replace s by s/k f That is. if the unsealed network has the network function H'(5), then
.
Thus if the unsealed network had a property, such as center frequency, when S =jl,
then the scaled network has this property at s = jk f This is clear from (15.35), which
.
406 Frequency Response Chap. 15
gives
HC/M = H'(yi)
Z'(5) = SV + R' + ^L
is any impedance in the unsealed network, then the corresponding impedance in the
scaled network is
Z(s) = sL +R J_
sC
= s iv) + R + '
57c7*7)
-I
=
R R' (15.36)
Therefore, to frequency-scale a network by a factor A7, we divide the L's and C's by k f
and leave the 7?'s unchanged. If there are dependent sources with constant gains, these
also are left unchanged.
To illustrate, the circuit of Fig. 15.4 is a low-pass filter with a cutoff frequency of
co c =
»J~2 rad/s, as we have previously seen. Suppose we want to frequency-scale the
network so that the cutoff frequency is 2 rad/s. Then the scale factor k f is given by
y/2k f = 2
or k f =
*J~2 Dividing the Cs in Fig. 15.4 by A7, we have, in the scaled network,
.
capacitances of lf2^/~2 F and 1/4^/T F. The rest of the circuit remains unchanged,
there being no inductors to scale.
Of course, we may perform both impedance and frequency scaling on a network.
To obtain a network with a practical property such as center or cutoff frequency, we
may first apply frequency scaling with the proper factor k f Then to make the resulting .
the parallel RLC network of Fig. I5.l contains a l-Q resistor, a 2-H inductor, and a
i-F capacitor, then for the network function of (1 5.5) we have an amplitude response
as in Fig. 15.2(a), with a resonant frequency ofrad/s and a peak amplitude of Ci.
I I
capacitance is
1/2 _ 1
10
k,k f ~ 2k X (
I0 6
._ 2k,-
,
10 '
H 1 mH
R I A, 500 fi
The scaled network is shown in Fig. 15.10(a), and its amplitude response, a scaled
version of Fig. 15.2(a), is shown in Fig. 15.10(b).
(o)
500--
(b)
EXERCISE
15.6.1 Frequency- and impedance-scale the circuit of Fig. 15.4 to obtain co c = 200071
rad/s (fc = 1000 Hz), using capacitors of 0.01 and 0.005 fiF.
Ans. k f = (lOOOv/Dn, k, = 10 5 \2J~1 n, R's = 22.5 kfi each
arbitrary scale) and it must be increased to l.l before the listener detects any change,
then the same listener can detect no change, if the original level is 2, until it is
increased to 2.2. In other words, the ear is not a linear device but more like a loga-
rithmic device, since the differences,
and
VzUco)
H(jco)\
V,Oco)|
Historically, the logarithmic unit, now known as the bel, was defined originally by
Alexander Graham Bell (1847-1922), who, of course, invented the telephone, as the
power unit,
bels = log 10
p
-^
However, this proved to be a large unit, so that the unit decibel (yL bel) became
common. That is.
dB 10 log
^
Chap. 15 Frequency Response 409
It" the two average powers />, and /', are referred to equal impedances, the last
expression ma\ be given in terms of the corresponding voltages by
dB 10 log,.
V2OC0)
V,Oo>)
which, of course, is equivalent to (15.37). In any case. (15.37) is taken as the standard
definition.
In practice, frequencies are not ideally blocked or attenuated'in the filtering process,
as may be seen in the low -pass response of Fig. 15.5(a). A zero amplitude would
correspond to absolute, or infinite, attenuation, and any approach to such an ideal
situation would be difficult to appreciate on a linear sketch. For example, if \W(j(x>)\
- 0.001 in Fig. 15.5(a) (jL of 1 ",, of its peak value of I), would correspond
this to
— 60 dB (or 60 dB below its peak value of dB). The latter figure means much more to
a telephone engineer than the linear figure.
Even more often in practice it is of interest to consider the attenuation or loss,
defined bv
a(co) = — 20 log , I
H( jco) I
-201og 10
y 2 Uco)
V,(»
V,(./co)
20 log 10 dB (15.38)
VzUco)
H(5)
s
2
+ J~ls + I
1
HOw)
VI +co 4
I
which is that of a low-pass filter with co c = 1 rad/s. The linear sketch looks somewhat
like that of Fig. 15.5(a). The attenuation is given in decibels by
a(<w) = 201og 10 J\ + w 4
- I01og, (l I- a/) (15.39)
I 2 3 4
EXERCISES
15.7.1 Let the amplitude response |H(y'cu)| be such that |H0'w)| max = |H(jO)| = K,
so that at cutoff |H(yco c )| =
K\«J~2 Find the loss, given by (15.38), at co =
'.
0.2s
His)
s2 + 0.2s + 1
find the loss in decibels at co = 0.0001,0.5,0.905, 1.0, 1.105, 10, and 100 rad/s.
Ans. 94, 18, 3, 0, 3, 34, 54
PROBLEMS
15.1 For the circuit shown, R = 2 Q, L =
H, and C = 0.01 F. If the input and1
PROBLEM 15.1
15.2 For the circuit shown, R= 1 ft, L= 0.01 H, and C = 1 F. If the input and
output are V, and V2 , respectively, find the network function and sketch the
amplitude and phase responses. Show that the peak amplitude and zero phase
occur at co = 10 rad/s.
PROBLEM 15.2
PROBLEM 15.3
412 Frequency Response Chap. 15
amplitude and phase responses. Show that the peak amplitude and zero phase
occur at co = 10 rad/s.
15.4 For the circuit shown, find the network function, H(.v) = V2 (.
s)[Vi(s )i and
sketch the amplitude and phase responses. Show that the peak amplitude and
zero phase occur at CO — 0.
PROBLEM 15.4
15.5 Show that in the general case the network function in the figure for Prob. 15.1
is given by
H(.s)
\ 2 (s) (R/L)s
V,(s) s2 + (R/L)s +
{l/LC)
Thus by comparison with ( 1 5.22) and ( 1 5.25), show that the circuit is a bandpass
filter with resonant, or center, frequency <o = l/s/LC, bandwidth B= R/L,
and quality factor Q = co /B - {\jR)^TjC. Show also that the gain K of the
filter, defined as its peak amplitude value, is given in this case by K = 1.
(MQ)s
H(j)
s2 + (llQ)s + 1
15.7 Show that in the general case the network function in the figure for Prob. I 5.2
is given by
\ 2 (s)_ (\/RC)s
H(.v)
V,(j) .9- +[(/? + \)IRC]s + {\/LC)
Thus show that the circuit is a bandpass filter with center frequency, bandwidth,
co
«JLC
R + I
B
RC
R
Q R 1 1
K
* + 1
3
(K/Q)s
H(.v)
s2 + (\IQ)s + 1
Thus the circuit lias resonant frequency co 1 rad/s, quality factor Q, gain
A. and bandwidth fi \!Q rad/s.
15.9 Show that in the general case of the figure for Prob. 15.3
V,(.v) (1/A,).v
H(.v)
V,(J) .v- :
<1/A\).v : (l//? 3 )
Thus show that the circuit is a bandpass filter with center frequency l/\/A 3
rad/s, bandwidth \jR z rad/s, gain R Z IR\, and quality factor /? 2 / v//? 3 Let .
A5 ,
= Q/K, A\ -= Q, and A\, = to obtain the network function of Prob.
1
15.8.
15.10 Using the results of Prob. 1 5.9, obtain a bandpass filter having the form of the
figure for Prob. 15.3 with co I rad/s, Q = 10, and A = 2. Find the approxi-
mate cutoff points and the bandwidth.
15.11 Show that the network function V 3 /V, in the figure for Prob. 15.1 is given by
V 3 \ILC
V, s2 - (RIDs + (1/Z.C)
15.12 Show that the circuit of Prob. 15.4 is a low-pass filter with a) c
= 1 rad/s. Com-
pare its amplitude response with that of Prob. 15.11 and note the improvement
around the cutoff point.
15.13 One type of low-pass filter is the //th-order Butterworth filter, whose amplitude
response is
\HUa>)\ =-
rT
J^, n = 1,2,3,...
1/A,
Hfr)-V.M_
V,U) s2 + (l/R 2 )s r (1/A 3 )
41 4 Frequency Response Chap. 15
|H(jw)|
PROBLEM 15.13
V*_ s^
V, s2 + (R/L)s + (XjLC)
15.17 Apply impedance and frequency scaling to the circuit of Prob. 15.1 to obtain
a bandpass filter with a center frequency f = 10 5 Hz, with a quality factor
6= 5, and using a capacitor of 1 nF. (Suggestion: See Prob. 15.6.)
15.18 Obtain an active bandpass filter using the configuration of Prob. 15.3 with
f = 1000 Hz, 6 = 10, and K= 2, using capacitors of 0.01 //F. (Suggestion:
See Prob. 15.9.)
K{s' I)
Him
V, (i/t».v i
rad/s. Also, as in the bandpass case, is the quality factor, and B (o„IQ \jQ
is the bandwidth. Note that the gain is H(0) A', where K 1
l-K
K
Q(l-K)
© I
Q(l-K)
J_
O
PROBLEM 15.19
15.20 Scale the network in the figure for Proh. I 5. 1 9 so that co = 10 6 rad/s, Q 5,
15.21 Show that the network of Prob. 11.18 is a band-reject filter with a gain of I,
15.22 The given circuit is a fourth-order bandpass filter (the denominator of the
network function is fourth degree), with co = */2 rad/s and = 5. Verify
this by finding
25'
H(.s)
N -i 2f3 T 102
r
5 5
and so forth. [Suggestion: Find the amplitude response and verify that
o> = peak point and that (D C[ C =& co =p (a> /10) satisfy the
-y/2" yields the ,
„
occurs when
^|H(/O»p=0
Since (d/dco) \
H |
2 = (d/da> 2 ) \
H |
2
(d(o 2 ld(o) = 2o> (d/da) 2 ) |
H =
|
2
0, we
must check co = separately.]
PROBLEM 15.22
16
TRANSFORMERS
416
; 7
Chap. 16 Transformers 41
In See. 7.4 we found that the inductance /. of a linear inductor, as shown in Fig. 16. 1,
A = N<f> = Li
dk j di
dt dt
It is evident from Faraday's law that a voltage is induced in a coil which contains
a time-varying magnetic flux, regardless of the source of the flux. Let us therefore
consider a second coil, having N z turns, positioned in the neighborhood of the first
coil, having N=N x
turns, as shown in Fig. 16.2(a). In this case, we have formed a
simple transformer having two pairs of terminals in which coil 1 is referred to as the
primary winding and coil 2 as the secondary winding.
To introduce several important inductive quantities, let us begin by examining the
open-circuited secondary case of Fig. 16.2(a). The current /', produces a magnetic
flux 0,, given by
where LI is the flux of/', that links coil 1 and not coil 2. called the leakage flux, and
;
is the flux of/', that links coil 2 and coil 1, called the mutual flux.
We shall assume, as in the case of the linear inductor, that the flux in each coil
links all turns of the coil. Since the secondary is an open circuit, no current flows in
A2 = ^ : 21
v 9 (open)
(a)
(c)
418
— 9
Chap. 16 Transformers 41
,.
• *
_ dX
— —2 -. -— /V N —dt
d<f> 2
;
,
2
~
dt
/V,0 : ,
M Z J, (1 6. 1)
'"'
dX t
St
where X t ,
the flux linkage of coil I. is gi\en by
A, -
A/,0,,
As in the case of a single linear inductor, the inductance L, of the primary winding,
sometimes called the self-inductance of the primary to distinguish it from the mutual
inductance, is defined by
, di ,
'"' L
'7/T
where (j) l2 is the leakage flux of/', that links coil 2 but not coil I. and i: is the mutual
flux of/' : that links coil 1 and coil 2. The flux linkage of coil I is
A, = yv,0 12
(//, ., dd>,,
dt dt
420 Transformers Chap. 16
If we define
Nrf 12 = M X2 i 2 (I6.3)
In the next section we shall show that the mutual inductances Af 12 and M 2X are
equal; therefore we shall write
M=M =M 12 lx (16.4)
and refer to M
as the mutual inductance.
The secondary voltage is given by
dX^
V2 = —J
dt
where the flux linkage is
X2 =N 2 <f> Z2
LJ 2 = N 2 <t> 22 (16.5)
2 ~
dt
respectively. Therefore the flux linkages of the primary and secondary coils are
X x
- A/,0 n 4- N i <p l2
X2 = N 2{ + N
2 (f> 2 (f> 22
Substituting from (16.1)-(16.5), we find upon differentiation that the primary and
secondary voltages are
V, L x ^ + M dt
(16-6)
2
dt dt
Chap. 16 Transformers 421
h is clear that the voltages consist of self-induced voltages due to the inductances / ,
and /.. and mutual voltages due to the mutual inductance A7.
In the preceding discussion, the coil windings in Fig. 1 6.2 are such that the algebraic
sign o\' the mutual voltage terms \l <//, dt and M di 2 dt arc positive for the terminal
voltage and current assignments as shown. In practice, it is, of course, undesirable to
show a detailed sketch of the windings. This is avoided by the use of a dot convention
which designates the polarity of the mutual voltage. Equivalent circuit symbols lor
the transformer o\~ Fig. 6.2 are shown in Fig. 6.3. The polarity markings are assigned
1 1
'2
(a) (b)
FIG U R E 1 6.3 Circuit symbols for the transformer of Fig. 16.2: (a) dots
on upper terminals; (b) dots on lower terminals
For the purpose of writing the describing equations, we may state the following
rule:
We note that with this rule it is unimportant whether the current i is increasing or
not. since the sign of the induced voltage is accounted for by di/dt. That is, if /'
is
increasing, the induced voltage M di/dt is positive, and if / is decreasing, the induced
voltage is negative. If / is a dc current, then, of course, the induced voltage is zero.
As an example, let us write the loop equations for Fig. 16.3(a). We see that /,
enters a dotted terminal. Thus the mutual voltage M di 2 /dt has positive polarity at
the dotted terminal of the primary. Similarly, since /', enters a dotted terminal, the
mutual voltage M dijdt has positive polarity at the dotted terminal of the secondary.
Application of KVL around the primary and secondary circuits gives (16.6).
In a like manner, identical statements apply to the undotted terminals of Fig.
16.3(b), and KVL once again yields (16.6).
Let us now establish a method for placing polarity markings on a transformer.
We begin by arbitrarily assigning a dot to a terminal, such as terminal a of Fig. 16.4(a).
422 Transformers Chap. 16
(a) (b)
polarity of the induced voltage, is a consequence of a rule known as Lenz's law. The
reader may study this law, named for the German scientist Heinrich F. E. Lenz, in a
later course in electromagnetic field theory.) We note that a current flowing into b
in Fig. 16.4(a) flux in the same direction as 0. Thus terminal b receives the
produces a
polarity dot. The symbol for this transformer is shown in Fig. 16.4(b).
circuit
Evidently the polarity markings on a transformer are independent of the terminal
voltage and current assignments. Consider, for instance, the assignment of Fig. 16.5(a).
Applying KVL around the primary and secondary loops, we find
/.
dt
M clL
dt
di,
M didt1 + L 2
dt
In Fig. 16.5(b), the voltage and current assignments have been changed. Application
of KVL for these assignments yields
(a) (b)
'
i
— L,—r- — M—r
dt dt
v2
,di
M^i
~d7
+ {
,
L^
,
Ll
di :
~dt
The above results can also be obtained by an alternative method for selecting the
sign of the mutual voltages, as follows:
If both currents enter (or leave) the dotted terminals of the coils, the mutual
and self-inductance terms for each terminal pair have the same sign; otherwise,
they have opposite signs.
The student can easily verify the use of this method for the cases of Figs. 6.3 and 6.5. 1 1
As an example, let us find the open-circuit voltage v 2 in the circuit of Fig. 16.6,
given that i\(0~) = 0. For / > 0, since no current flows in the secondary, applying
KVL around the primary yields
/, = 2 + Ae- i0!
i, = 2(1 - e~ ]0:
) A
Therefore, in the secondary.
v2 = -0.2541
dt
= -5e- 'V 10
Let us now consider the transformer of Fig. 16.3(a) in the case of an excitation
having a complex frequency s. In this case, the phasor equations corresponding to
(16.6) are
Y l
(s) = sL 1
l l (s) + sMU(s)
V z (s) = sMl^s) + sL 2 l 2 (s)
The phasor circuit for this network is shown in Fig. 16.7. In the case of a purely
sinusoidal excitation, we simply replace s by jco.
For a final example, consider finding the phasor voltage V 2 (5) in the network of
Fig. 16.8 due to the complex forcing function V,(s). The loop equations in the primary
424 Transformers Chap. 16
0.25 H sM
20 V^ V, Sl_, Sl_,
FIGURE 16.6 Circuit with an open-circuit secondary FIGURE 16.7 Phasor circuit for a
transformer
Therefore
s + 2 V,
r
s
l
2
s + 2 s
s 2s + 3
-*v,
s
2
+ Is + 6
_2 31^ -3s
H(s)
V, s
2
+ Is -
-35
(s + 1)0? + 6)
A pole-zero plot and a sketch of |H(/to)| are shown in Fig. 16.9. We may show that
the maximum ac steady-state response occurs for co = */~6 rad/s.
Suppose, as an example, that t>, = 100 cos 10? V. Then
I
co
-6
*— 9 a
(a)
and
V, = 25.6 /126.7° V
EXERCISES
16.1.1 In the circuit of Fig. 16.3(a), L = L =0.1 H and A/ - lOmH. Find r, and
x 2
v 2 if (a) /, =10 mA and i2 = 0, (b) = and i2 = 10 sin 100/ mA, and (c)
/,
16.1.2 Verify that the coil windings of Fig. 16.2 are consistent with the polarity
markings of Fig. 16.3.
»(/) = \Li\i)
Evidently, for a given inductance L. the energy is completely specified in terms of /'(/)
426 Transformers Chap. 16
Let us now determine the energy stored in a pair of mutually coupled inductors, such
as those of Figs. 16.2-16.4. In so doing, we shall show that A/ 12 = M 2l = M and
establish limits for the magnitude of M.
In Fig. 16.3, the stored energy is the sum of the energies supplied to the primary
and secondary terminals. The instantaneous powers delivered to these terminals,
from (16.6), are
Pi = v iH =
\
M ^-^ + ^"jfV*
where M has been replaced by M X2 and M 2i in the appropriate terms.
Let us now perform a simple experiment. Suppose we start at time t with /,(f )
— '2(^0) — 0- Since the magnetic flux is zero, no energy is stored in the magnetic
field; that is, w{t ) = 0. Next, assume, beginning at time t , that we maintain i2 =
and increase /, until, at time and i 2 {t ) = 0. During
/,, /,(?,) = /, x
this interval, i2 =
and dijdt = 0. Thus the energy accumulated during this time is
f" f" di
if, (Pi +Pi)dt= M,j<fi
During this time interval, dijdt = 0, and the energy accumulated, using (16.7), is
Wi Hm^ + l^Y
-I2
= (M 12 /, + L 2 i
2 )di 2
= M X2 IJ2 + \L 2 1\
Let us now repeat our experiment but reverse the order in which we increase/, and
i2 . That is, in the interval from t to t u we increase /, so that i 2 {t x ) = I 2 while holding
M 12 = A/ 21 = M
If we repeat our experiment with the polarity markings of Fig. 16.4, the sign of
the mutual voltage is negative, which causes the mutual term in the energy M/,/ : to
be negative. Thus a general expression for the energy at any time / is given by
\\ here the sign in the mutual term is positive if both currents enter dotted (or undotted)
terminals; otherwise, it is negative.
The coefficient of coupling between the inductors indicates the amount of coupling
and is defined by
k = <l6 9)
7lk -
M = JM, M
V0 M V0 22 -<
k = 1 1 ,
= fat I
/02_,
,
V/L,L 2 JL,L 2
0< k< 1
or, equivalently.
0< M <^L X
L2
If A- == 1, all of the flux links all of the turns of both windings, which is a unity-coupled
transformer.
The value of A (and hence M ) depends on the physical dimensions and number of
turns of each coil, their relative positions to one another, and the magnetic properties
of the core on which they are wound. Coils are said to be loosely coupled if A < 0.5,
whereas those in which A > 0.5 are tightly coupled. Most air-core transformers are
loosely coupled, in contrast to iron-core devices for which A can approach I.
Let us now examine the values of i 1 and z\ in (16.8) for which w(t) is zero. From the
quadratic formula, we may write
/, = T j^ ± f-JW- - L L l
X Z
For real values of i1 and /\. we see that for w(t) = we must have
428 Transformers Chap. 16
0, M < JL L X 2
M.
±—'2, M = ^/TJT
(16.10)
2
where the sign of the mutual term is negative if both currents enter dotted (or undotted)
terminals; otherwise, it is positive. The second equation of (16.10) shows that w(t)
can be zero in a unity-coupled (k = 1) transformer even though /, and i 2 are nonzero.
EXERCISE
»,(0 = 7?,/, I
L,
dt
M di.
M^\R 2i 2
1 di 2
Chap. 16 Transformers 429
Evidently, the voltages and currents are not affected byV For this reason, the secon- .
we now let V = 0. it is seen that the bottom terminals of the transformer are con-
nected and that the primary and secondary circuits have a common reference point.
The transformer, in this case, is a three-terminal device.
As a second example, let us find the complete response /, for / in Fig. 16. 1 1,
2a
given M=
\\*f~2 and /,(()") / 2 (0~) H =
0. Since the forced response for /, is a dc =
current (by inspection), no steady-state voltage is produced in the secondary; thus
i
2f
= 0. To find the natural response, let us first obtain the network function for / > 0.
3 „\ ¥
V,(5) = (-|j +
,
2)l , - -jjsl
l
1
= sh (s + 2)I 2
H(J) = I
^ 2 (s + l)(s + 4)
The poles of H(s) are the natural frequencies. — 1, —4, of the natural response.
Thus, for / > 0.
h = 1
2/ i iH = +A 1
e-' + Aie-
From (16.8), w{0~) = 0; therefore since the energy cannot change instantaneously
in the absence of infinite forcing functions, w(0 + ) = 0. From (16.10), noting that
M< JL L 1
A we1 ,
+
see that /](0 ) = = To obtain a second initial condition for
/ 2 ( u+ ) u -
yields
3 J/ 1 (0 + )
+
12 = 1 di 2 (0 )
2 dt dt
+ +
rf/,(0 ) di 2 (0
= 1
)
dt dt
430 Transformers Chap. 16
KS}
5s + 4
and the complete response is
i2 = Ae-°*'
I? - ?i -4-
3 dl '
/
3 di 2
0= _7Z^l
V2(/f +
+^
f 2/,2t A
Multiplying the latter equation by v/3/2 and adding it to the former, we have
Recalling (16.10), we see that the energy is zero in the unity-coupled case when
/,(0
+
) = ^/ 2
(0+) = ^i 2 (0 )
+
i2 = 2.94c- - 8 '
°- -( s + t) i + '
(
2i + 1+ t)
for which
H(s)
V _ h. -2
S2 + l
V, V! s3 + 3s
2
+ 2s + 2
Chap. 16 Transformers 431
v, = 10 cos 2t
As a final us find the network function V 2 /V, for the phasor circuit
example, let
of Fig. 16.13. Let us examine the voltage V^ which appears across winding A in
first
loop 1. We see that the voltage due to I, is the self-inductance term (2.vl ) and the ,
mutual term of winding C (2sl,). The voltage due to I 2 arises from the mutual terms
of winding B — .d,) and winding C — 2.si 2 ). Thus
( (
Similarly, we find
0= -21, - V c + \ B +5\ 2
or
V,= (lO.v + 3)1, - (10s L 2)I 2
Solving these equations for the ratio I 2 /V,, we find the network function to be
H(j)-
3I 2 3(10j+ 2)
V, ~ V, ~
30j 2 + 49^+11
EXERCISES
16.3.1 Determine i l
for / > in the network of Fig. 16.11, given M = l/*fT H.
Assume the circuit is in steady state at / = 0~. Ans. 6 — 4e~' — 2e~ 4! A
16.3.2 Determine the impedance seen from the terminals of the voltage source v\ in
Fig. 16.12. Ans. 1.48 /37.75 Q.
16.3.3 Determine the network function V 2 /V, for Fig. 16.13 if the polarity dot of coil
B is placed on the other terminal. Ans. 6(5 + l)/(6s 2 + 45^ + 11)
In this section we shall develop several important impedance relationships for the ac
steady-state case. Let us begin by considering the phasor circuit of Fig. 16.14 having
a practical source \s with an impedance Z, connected in the secondary.
L
2
V2
D
l2 I
V, =jcoL 1 l 1
-jcoMl 2
(16.12)
= -jcoMI l + (Z 2 +jo>L 2 )l 2
Eliminating I
2
from these equations we have
jcoMijcoM)'
V, JcoL ">
1
Z 2 + jcoL 2 _
—L
Chap. 16 Transformers 433
so that the input impedance seen at the primary terminals of the transformer is
Z == : -; Q,L '+ 16 13 )
'
l7 Z 2 +./o>L 2 ( -
The first part. jcoL^ depends entirely on the reactance of the primary. The second
part is due to the mutual coupling, and it is called the reflected impedance, given by
7 or A/ 2
£jR
-Z 2 +jcoL 2
It may be thought of as the impedance inserted into, or reflected into, the primary by
the secondary.
The input impedance as seen by the source Vg is evidently
z in
= zg + z,
Also the secondary-to-primary current ratio I 2 /I, may be found from the second
equation of (16.12). and the voltage ratio V 2 /V, may be found from
V
T, ~Z >«)
using (16.13) The result is
jcoM
I, Z 2 |
jcoL 2
(16.14)
v 2
jcoMZ 2
v, ja>L (Z 2 x + jcoL 2)
-\- a, 2 M 2
z" = - jtt+<0li)l
5iTTO? [S!
ln*J 2 C
434 Transformers Chap. 16
EXERCISES
16.4.1 Given: In Fig. 16.14, \g = 100/0° V, Zg = 10 ft, L, = L2 = 0.5 H,
M= 0.1 H, and co = lOOrad/s. If Z2 = 10 - O'1000/o>)ft, find (a) Z in ,
Arts, (a) 48.81 /77.47° ft, (b) 2.05 / -77 .47° A, (c) 0.50 /-63.43 A,
(d) 97.62 /11.83° V, (e) 7.07 /- 108.43° V
16.4.2 Repeat Ex. 16.4.1 if the polarity dot is on the lower terminal of the secondary.
Ans. (a) 48.81 /77.47° ft, (b) 2.05 / -77.47° A, (c) 0.50 /116.57° A,
(d) 97.62 /11.83° V, (e) 7.07 /71.57° V
16.4.3 Determine the resonant frequency in Ex. 16.4.1 for which the reflected imped-
ance is real. Ans. 7.12 Hz
=
n
£ (16.15)
where N and N 2 are the number of turns on the primary and secondary, respectively.
x
proportional to the product of the current and the number of turns on the winding.
Thus, in the primary and secondary windings,
0ii = a/V,/,
<f>22 = aW 2
/
2
L, = a/V
2
, L2 = o.N\
Therefore
£-(&)'—* (16.16)
jcoZ 2 ^/L ]
L 2
= n
For the ideal transformer, in addition to unity coupling, the inductances L, and L 2
tend to infinity in such a way that the ratio of (16.16) is the constant n 1 . From the first
h = Joi^/L L2
ij m s
Ij l,.l,-~ Z 2 + jcoL 2
- lim
JQ>*JLJL 2
LuLr^jCO + (Z 2 /L 2 )
= ]im MV")
L,,L,—>j<Q + (Z 2 /L 2 )
J_
n
Thus the primary and secondary voltages and currents of an ideal transformer are
by
related simply to the turns ratio
v 2
rr- (16.17)
h _ l
n
436 Transformers Chap. 16
Replacing n by N 2 /N u we have
N 2l2
Therefore the voltages are in the same ratio as the turns, and the ampere turns (NI) are
the same for both primary and secondary.
The symbol for an ideal transformer is shown in Fig. 16.15, connected to a secon-
©9 |
•
V2
-
s^\
i')
Z
2
dary load Z2
and a source V g with impedance Z g The vertical lines are used to sym-
.
bolize the iron core,and 1 n denotes the turns ratio. If one or the other, but not both,
:
of the polarity dots is placed on the opposite terminal, then the ratios of (16.17)
become negative.
The primary impedance Z, of (16.13), in the case of the ideal transformer, is given
by
_ V, _ VJn _ V 2 /I 2
7
or
7 _Z 2 .
1 — »2 (16.18)
z,
where Z2 is the load impedance of Fig. 16.15. Thus the input impedance viewed from
the terminals of the voltage source is
Z in g
(16.19)
n2
If = n\%
from which
V 2 If
If
or
V,If _V 2 If
Chap. 16 Transformers 437
Thus the complex power applied to the primary is delivered to the load Z 2 , and hence
the transformer absorbs zero power.
In anal\ zing networks containing ideal transformers, it is often convenient to
remove the transformer before performing the analysis. Let us consider, for example,
replacing the transformer and load impedance Z 2 of Fig. 16.15. Clearly, the input
impedance seen by the generator V, is Z in given by (16.19), so that an equivalent circuit
insofar as V s is concerned is shown in Fig. 16.16. The voltages and currents can now
be easily determined from the single-loop circuit.
Zg
Z?/n V, =
Let us next consider replacing the primary circuit and the transformer of Fig.
16.15 by its Thevenin equivalent. By (16.17) we have
I, = „I 2 . V2 = A?V,
V oc = V 2 = #iV, = n\ g
For I sc , we have V2 = and thus V, = 0. Therefore
V.
T
*sc
sc
=
— I
*2
-*2
I,
—- ii t
and
th — j— - n ^ g
27
n Zg
Z2 V 2 =nV,
" = I,/n
p
FIG U R E 16.17 Thevenin equivalent circuit of Fig. 16.15
438 Transformers Chap. 16
Inspecting the results for the Thevenin equivalent circuit, we see that each primary
voltage is multiplied by n, each primary current is multiplied by \jn, and each primary
impedance is multiplied by n 1 when we replace the primary circuit and the transformer.
It may be shown that this statement holds in general whether the result is the Thevenin
circuit or not. On the other hand, if we replace the secondary circuit and transformer
by an equivalent circuit, as in Fig. 16.16, we simply multiply each secondary voltage,
current,and impedance by 1//?, n, and l//? 2 respectively. If either dot on the transform-
,
(a)
iooZO° 300+1400 a
(b)
V + ioo/o_°
2
+ io 3
v2 +
10 3 300 + /400
from which
V2 = 27.95/206.6° V
Chap. 16 Transformers 439
So far we have considered only the ideal transformer in the ac steady-state case.
In the general case, we see from (16.6) that
Ll
~dT \L 2 L2 dt )
Thus since M =L 2
l
L 2 and ^'LJL^ — /?, we have
t\ _ di ]
+
M di 2
L "~dJ
x
L[ dt
di, di\
=
^+
,
,1
-dT
/', = —n\ 2 + C,
where C x
is a constant of integration. Since dc currents produce no time-varying
magnetic flux, they do not contribute to the induced voltages or currents in the ideal
transformer. Therefore if we neglect the constant C,, then
i, = -ni2
where the minus sign arises due to the direction assigned i 2 in Fig. 16.3(a). Thus the
same current and voltage relationships are valid in the time domain as were found in
the frequency domain if we neglect any dc currents.
EXERCISES
16.5.1 In Fig. 16.15, V, = 100/0^ \,Z g = 100 ft, and Z2 = 90 kQ. Find // such that
Z, = Zg , and then find the power delivered to Z2 . Ans. 30, 12.5 W
440 Transformers Chap. 16
16.5.2 Using Norton's theorem, show that the primary circuit and transformer of
Fig. 16.15 is equivalent to a constant current source of V g /nZ g A in parallel
with an impedance of n 2 Z g fi.
16.5.3 Find V2 in Fig. 16.18 (a) by first replacing the secondary circuit and the trans-
former by an equivalent circuit.
Equivalent circuits for linear transformers are easily developed by considering the
equations for primary and secondary currents and voltages. In Fig. 16.19(a) we have
1
'
dt
~ dt
M^i + L
dt
2
^
di :
Tt
It is evident that the circuit of Fig. 16.19(b) satisfies these equations. The dependent
voltage sources, however, are controlled by the time derivatives of the primary and
secondary currents. In the frequency domain, these sources can be considered as
current-controlled voltage sources.
°—TiRP 1
—^WT^—
i
o
&<£> 6 dt
(o) (b)
idi, ,
,,/di, ,
dL
m( di, ,
di-
) + (i
>
- M >£
These equations are satisfied by the T network of Fig. 16.20(b). Since this circuit is
a three-terminal network, it is equivalent to the transformer connection of Fig.
16.20(a). If either polarity dot is changed to another terminal, we must replace by M
—M in the equivalent circuits.
Chap. 16 Transformers 441
o—nnnr - 1
+ L,-M L 2 -M +
(o) (b)
In the case of an ideal transformer as shown in Fig. 16.21(a), the currents and
voltages are given by
i2 = —
n
!
>
21
=
v2 nv,
Clearly, the circuits of Figs. 16.21(b) and (c) satisfy these relations. If either of the dots
are reversed, of course, we must replace n by —n in the equivalent circuits.
(a) (b)
'2
V ni nv V
l
2<J> <d> l
2
(c)
FIGURE 16.21 (a) Ideal transformer; (b) and (c) equivalent circuits
442 Transformers Chap. 16
EXERCISE
16.6.1 Employing the T equivalent circuit for the linear transformer in Fig. 16.12,
determine the phasor current Ij. Ans. 6.11 /— 31.15° A
PROBLEMS
16.1 In Fig. 16.2 (a), 7V2 = 1000 turns and <f> 2l =50 ^Wb when /, = 1 A. Deter-
mine v 2 if i\ = 10 cos 100/ A.
-oc
k><>
(a) (b)
PROBLEM 16.2
16.3 In Fig. (a) of Prob. 16.2 found that the inductance measured between ter-
it is
PROBLEM 16.4
0.5 H
I A< v
PROBLEM 16.6
16.7 Repeat Prob. 16.6 if a 1-F capacitor is connected in parallel with the \-Q
secondary resistor.
16.9 Substituting (16.9) into (16.8), show that a transformer satisfies the passivity
condition w(t) > 0.
16.10 Repeat Prob. 16.4 if a 3-Q resistor is connected to the secondary terminals.
(Hint: Replace the current source and 2-fi resistor by their Thevenin equiva-
lent.)
16.11 Repeat Prob. 16.4 if the mutual inductance is «/2 H and a 3-Q resistor is
I3A 4D
47 cos 1000
V 6 2mH ImH 3mH
PROBLEM 16.12
i(t)
2H
v(t) Z^±F 4H
VW^
en
PROBLEM 16.13
444 Transformers Chap. 16
I.(s)
OIF
in
V(s)
6 2H 5H
PROBLEM 16.14
IA
2
H IH m
H
v,(s)r + IH 2
h ia< v,( S )
PROBLEM 16.15
F
I" 1:2
i
cos 2t
V © 2n
PROBLEM 16.16
PROBLEM 16.17
'
AA/*
2A
100 V
rms © 4n
PROBLEM 16.18
16.19 In Fig. 16.20 (a), determine the r-parameters for the transformer for complex
frequency s.
VvV
10X1
20 cos 100 t (
V
PROBLEM 16.21
.
17
FOURIER METHODS
Thus far, with few exceptions, we have considered excitations which were exponentials,
sinusoids, or damped sinusoids, and the reader may have been impressed by the
relative ease with which we can handle such functions. In particular, the concept of
impedance allowed us to treat circuits with these excitations as if they were resistive
circuits and find both their forced and natural responses.
However, as we all know, there are many functions that are important in engi-
neering which are not exponentials or sinusoids. A few examples are square waves,
sawtooth waves, and triangular pulses. Indeed, a function may be represented by a set
of data points and have no analytical representation given at all. In this chapter we
shall consider these additional functions and show how we may represent them in
terms of our familiar sinusoidal functions. The techniques we shall use were first given
in 1822 by the great French mathematician and physicist Jean-Baptiste Joseph
Fourier, who lived from 1768 to 1830.
Let us begin by considering a function f(t) which is periodic of period T; that is,
f(t)=f(t+T)
f(t) = -j + a x
cos co t + a 2 cos 2co t + . .
446
7
fit) = ^r
Z
+ 2 ( fl n cos nc<V + *n sin nco t) (17.1)
n - 1
where co = 2^/r. This series is called the trigonometric Fourier series, or simply the
Fourier series. of/(/). The as and 6's are called the Fourier coefficients and depend, of
course, on/(r).
The coefficients may be determined rather easily by the use of Table 12.1, given
earlier in Chapter 12. Let us begin by obtaining a , which may be done by integrating
both sides of (17.1) over a full period; that is,
•r t>T
I fif)dt=. (^-dt
Jo Jo L
Since 7" = 2n/co , every term in the summation is zero by entry 2 in Table 12.1 (a = 0),
and therefore we have
°o =^
r.f fit)dt (17.2)
Next, let us multiply (17.1) through by cos mco Q t, where m is an integer, and
integrate. This yields
+ Y^bn \
cos /mgv sin nco t dt
«=i Jo
By entries 2, 4, and 5 of Table 12.1 (for a = /? = 0), every term in the right member is
zero except the term where n =m in the first summation. This term is given by
T rp
am I cos
2
tncoct dt = — a m = -^a„
o ^o
t»n
so that
Finally, multiplying (17.1) by sin ww r, integrating, and applying Table 12.1, we have
We note that (17.2) is the special case, m = 0, of (17. 3) (which is why we used
a /2 instead of a for the constant term). Also, as the reader may easily show, we may
integrate over any interval of length T, such as t to t T, for arbitrary t and the + ,
results will be the same. Therefore we may summarize by giving the Fourier coefficients
in the form
['°
+T
_2
T fit) cos nm Q t dt, n = 0,1,2,
(17.5)
In any finite interval /(0 has at most a finite number of finite discontinuities,
a finite number of maxima and minima, and
\
\f(t) dt < co (17.6)
Jo
37T t
Since T 2n, we have eo = 2n/T = l . If we choose t = —n, then the first equation
of (17. 5) for n = yields
a = — I C"
t dt =
For /; 1. 2. 3, . . . , we have
t cos »* <//
n
~ n 2 n (cos nt + n* sin »0
=
and
2 COS «7I
«
2(-l)" +1
The case /? = had to be considered separately because of the appearance of/; 2 in the
denominator in the general case.
From our results the Fourier series for (17.7) is
/(,) = 2 (™i _ ^* + ™* _ . . .
)
(17.8)
The fundamental and the second, third, and fifth harmonics are sketched for one
period in Fig. 17.2. If a sufficient number of terms in (17.8) are taken, the series can be
made to approximate /(r) very nearly. For example, the first 10 harmonics are added
and the result sketched in Fig. 17.3.
As another example, suppose we have
fit) = 0, -2<t<- -1
= 6, -Kt< 1
values. If t I, we only have to di\ ide the integral into two parts since/(f) 6 on
I to I and /(f) on I to 3. Therefore let us choose this value of/ and obtain
Also we ha\e
2 I , nnt ,, 2 f ft /;7r/ .
12 . im
sin .
tm 2
and. finally.
3
2 f nt
/77T7 2 n /77H
/77W
,
o„ = -r
£
6 sin -=-
2" <at
,,
/?
,.
+
-•
,
,
-r
"4
f
°
..
s,n
sin —=- at
"2"
,
4 ' - 1
*- j i
T
.
-- 1
T cos- 5 -
. ]
T co St - -...)\
Since there are no even harmonics, we may put the result in the more compact form
+1
/(,) = 3 + 11 v (-l)" cos{[(2;7-
—
\)nt}/2}
71 „= i 2/7 1
EXERCISES
17.1.1 Find the Fourier series representation of the rectangular wave
f{t) = 3, 0</<l
= -1, 1 < t <4
/(' + 4)=/(0
^/7i-. a = 0, «„ = —
/77T
sin -=-, o„
2 ' "
= —
nn
I
V
1 — cos -=-
2 /
, n = 1, 2, 3, . .
452 Fourier Methods Chap. 17
= 0, \<t<n
fit + n) = fit)
Ans . l + ± ± sin 2(211-1)/
—
n „=
I
2« 1
/(0 = 0, -1 </< -i
4, -1 < / < I
= 0, } < <
f 1
If a function has symmetry about the vertical axis or the origin, then the computation
of the Fourier coefficients may be greatly facilitated. A function fit) which is sym-
metrical about the vertical axis is said to be an even function and has the property
f(r)=f(-t) (17.9)
for all /. That is, we may replace / by — / without changing the function. Examples of
1
even functions are t and cos f, and a typical even function is shown in Fig. 17.4(a).
Evidently we could fold the figure along the vertical axis, and the two portions of the
graph would coincide.
A function /(/) which is symmetrical about the origin is said to be an odd function
and has the property
/(')= -/(-/) (17.10)
In other words, replacing t by — tchanges only the sign of the function. A typical odd
function is shown in Fig. 17.4(b), and other examples are / and sin /. Evidently we can
fold the right half of the figure of an odd function and then rotate it about the /-axis
so that it will coincide with the left half.
Now let us see how symmetry properties can help us in determining the Fourier
coefficients. Evidently, in Fig. 17.4(a) we may see by inspection that for /(/) to be an
even function we have
f
" f{t) dt = 2 \"fit) dt (17.11)
J-o JO
Chap. 17 Fourier Methods 453
f(t)
(o) (b)
This is true because the area from —a to is identical to that from to a. This result
may also be established analytically by writing
= - Cf(-T)dT+ Cf«)dt
Ja Jo
= \'f(T)dt+ \"f{t)dt
Jo Jo
a
= 2 f{t)dt
Jo
\
In the case of/(?) an odd function, it is clear from Fig. 17.4(b) that
f f(t)dt = (17.12)
J -a
This is true because the area from —a to is precisely the negative of that from to a.
This result may also be obtained analytically, as was done for even functions.
In the case of the Fourier coefficients we need to integrate the functions
g(-t) =/(-Ocos(-to»oO
= f(t) cos nco t
= g(t)
. 7
and
K—0 =f(-0 sin (-nco t)
= -hit)
Thus g(() is even and h(t) is odd. Therefore taking t = — 772 in (17.5), we have, for
fit) even,
•772
/z
a„ = 4
-= r
\ f(t) cos nay <#, n = 0,1,2, ...
* Jo (17.15)
6„ = 0, n = 1,2, 3, ...
an = 0, n = 0,1,2,...
2
(17.16)
6„ = —4 f<
/(/) sin nco t dt n = 1, 2, 3, . .
In either case one entire set of coefficients is zero, and the other set is obtained by
taking twice the integral over half the period, as described by (17.15) and (17.16).
In summary, an even function has no sine terms, and an odd function has no
constant or cosine terms in its Fourier series. As examples, the functions given earlier
in (17.7) and Ex. 17.1.1 are odd functions, and their Fourier series have only sine terms.
The function of Ex. 17.1 .4 is even so that its series contains only cosine terms (including
the dc case, n = 0). The function of Ex. 17.1.2 is neither even nor odd, and conse-
quently both types of terms are present.
To illustrate the procedure, let us find the Fourier coefficients of
<'„ = 0, n = 0,1,2,...
= -5-P -(-1)"]
Therefore b„ = for n even and b n = 1 6/nn for n odd, as indicated in Ex. 17.1.1.
It often happens that the function /(/) is not periodic, and we shall consider an
alternative method later for this case. However, as is often the case, we may be
n
interested only in /(f) on some finite interval (0, T). in which case we can consider it as
periodic of period T. and find its Fourier series as before. Of course, what we have is
not the Fourier series of /(/) but of its periodic extension. We may, however, use the
result on the interval of interest and not be concerned with its periodic behavior
elsewhere. Alternatively, we may consider (0, T) to be halfxhe period and expand /(/)
as either an even or an odd function using (17. 15) or (17.16). The result, called a half-
range sine or cosine series, is easier to obtain and is equally valid on the interval of
interest.
EXERCISES
17.2.1 Find the Fourier series for the function of (17.7) using symmetry properties.
17.2.2 Find the Fourier series for the half-wave rectified sinusoid
/(/) = 0,
^ l
2 4
= 4 cos 2t,
4 4
= 0, ""
4 2
/(f + *)=/(/)
This function, shown in the accompanying figure, is a sinusoid with its nega-
tive amplitudes suppressed. Thus it is the output of a circuit that changes ac to
pulsating dc.
im
Am. a = — ;a, = 2;a„ =— 8
71(1 — n
=-\
2
)
cos -t> "
2
'
_ , A
= 2,'
3, 4,
' '
0, n = 1,2, 3, ..
4 2 2 4
EXERCISE 17.2.2
= +
17.2.3 Find the Fourier series for the full-wave rectified sinusoid
/(/) = |4sin2/|
This function, shown in the accompanying figure, is a sinusoid with its nega-
tive amplitudes changed in sign.
Ansl
4{\ + %T^W C0 * An*)
EXERCISE 17.2.3
Replacing the sinusoidal terms in the trigonometric Fourier series by their exponential
equivalents,
and
[(£^Ay-.. + (^±j^y
Kt)
f i (17.17)
an - jb n
(17.18)
and then substitute for a n and b n from (17.5), with t = — T/2, we have
Chap. 1 7 Fourier Methods 457
T 2
H r i
f(t)(cos nco t — j sin nco Q t) dt
if f(t)e-
Jna «
dt (17.19)
r* _ an + jb„
2
= i r 2
J-r/2
f =4 jTy»*
which by (17.19) is
^=c (17.21)
Summing up, (17.18), (17.20), and (17.21) enable us to write (17.17) in the form
/(/) = c + f] c n e
in "°'
+ f] c-jr
ia»*
= /BO
I] cne "" + 2 c„e
;ncu °'
We have combined c with the first summation and replaced the dummy summation
index n by —n in the second summation. The result is more compactly written as
where c n is given by (17.19). This version of the Fourier series is called the exponential
Fourier series and is generally easier to obtain because only one set of coefficients
needs to be evaluated.
As an example, let us find the exponential series for the rectangular wave of Ex.
17.1.1, given by
458 Fourier Methods Chap. 17
c„ = ij_ f(f)e-""dt
For a? ^ this is
c„ = A
f
(-4)e--""" cfr + A f
4<r-""
1
'
<#
J- 1 Jo
= 4-[i -(-i)i
Also we have
dt
r i
4rf 4 dt
= ir, '-il'
Since c„ = for « even and c„ = 8//'«7i for n odd, we may write the exponential
series in the form
f{t) = ~ t ^—re^"-^'
JTl n^=-° in — 1
The reader may verify that this result is equivalent to that obtained in Ex. 17.1.1.
EXERCISES
17.3.1 Find the exponential Fourier series for
f{t) = \, -\<t<\
= 0, 1 < 1
1 1
<2
fit + 4) = /(/)
Ans l+1 v sin (fflt/2) /2
n*0
c = lim c„
17.3.3 Generalize the results of Ex. 17.3.2 by finding the exponential series of a train
of pulses of width 5 < T, described by
/(0 = 1, -|-</<4
=0, 4<i,i<-£
f(t + T)=f(t)
Ans.± ^Sa(^)^-/r
We are now in a position to find the forced response of a circuit to any periodic
excitation that has a Fourier series. The excitation is expressed as a Fourier series,
and the response to each term of the series may be found by the phasor method. Then
the total response is, by superposition, the sum of the individual responses.
To illustrate, suppose we have an RL series circuit, with R = 6 Q and L = 2 H,
excited by a source
<t) = ±v n
where
°-
= (2„' 6 sin(2 '| -' ) '"
l)*
is the (2// — l)th harmonic of v(t), with frequency co„ = (2n — \)n rad/s.
The phasor voltage is
V = — /-90°
I. = |I,| /-90°-g.
where
o
II = -
(17 231
1
"'
|
and
/n (O = |I„|cos(aV-90 -0„)
= |I„| sin(ov ~ On)
M»=i;|l,,|sin(av-0 n) (17.25)
n= 1
We may obtain the natural response, and thus the complete response, as we have
done in the past. A difficulty with the procedure is the necessity for finding the arbi-
trary constant in the natural response by summing an infinite series. This, of course,
also presents difficulties in finding Very often, however, a reasonable
i(t) for a given t.
number of terms can be used, with the remaining terms ignored, to approximate the
series.
The input v(t) in the example we have just considered contains not just one
frequency, as was the case for sinusoidal excitations, but an infinite number of
frequencies. This is, of course, also true of the output /(/). Since the Fourier series
displays the amplitude, as well as the phase, corresponding to each frequency, it is a
simple matter to which frequencies are playing an important role in the shape of
tell
the output and which are not. This, of course, is one of the important applications of
Fourier series and will be considered in more detail in the next section.
Chap. 17 Fourier Methods 461
EXERCISE
17.4.1 Obtain (17.25) using the exponential Fourier series. [Suggestion: Recall that
We may write the /?th harmonic in the trigonometric Fourier series in the form of a
single sinusoid.
with amplitude
A„ = Jal + bl (17.27)
and phase
0„=-tan->^ (17.28)
_ a„ - jb n
c. = ^ fl
-+*-/-tan-^
2 / an
Therefore we have
cn = 4?/<f>„ (17.29)
An output function /(/), given in the time domain, may be represented in the
frequency domain by A n and 0„, n = 0, 1, 2, For example, if we wish to know if
. . . .
readily obtained from a plot of A„ versus frequency. Since the frequencies are discrete
values such as 0, co 2<w etc.. the plot will be a set of lines whose lengths are pro-
. ,
portional to A„. Such a plot is called a discrete amplitude spectrum or a line spectrum
and is analogous to the amplitude response considered in Chapter 15 for the con-
tinuous case. A similar line plot of 0„ is a discrete phase spectrum and is analogous to
the phase response in the continuous case. The dc component, of course, is given by
A = a /2 and = 0, or if a <0, A = a„/2and O = 180°.
<f> {) —
462 Fourier Methods Chap. 17
As an example, the rectangular pulse of Ex. 17.3.1 had Fourier coefficients given by
nn
1
nn l
n= ±1,±2, ±3,...
Also, from Ex. 17.3.2, we may write this result in terms of the sampling function as
sin (nnjl)
A = 2\c n \
nnjl
.-r^rvrvTV V IL VTVT\-TV-r-
5-3-10 1 3 5 7 n(co = n7772)
EXERCISES
17.5.1 Find the amplitude and phase spectra A„ and (f)„ of the sawtooth wave of (17.7).
Ans.A n =\2/n\,n = ±1, ±2, ±3,...
0„ = -90°, n = 1, 3, 5, -2, -4, -6, . . . , . . .
/(f) = 1, -a<t<a
= 0, a < | / 1 < 772
f(t + T)=f(t)
Chap. 17 Fourier Methods 463
where 772 > a. Note that this function is a generalization of that of Ex.
17.3.1 and that its spectrum envelope reaches zero at \/a, 2/a, 3/a, which . . . ,
to be periodic with an infinite period and extend our previous results to include this
case. In this section we shall consider this case in a nonrigorous way, but the results
may be obtained rigorously if fit) satisfies the conditions of Sec. 17. 1 with (17.6)
replaced by
Let us begin with the exponential series for a function fT (t) defined to be/(/) for
-T/2 < <t T/2. The result is
where
c„ = -^ fT {x)e-
iUnx/T
dx (17.32)
We have replaced co by 2^/Tand are using the dummy variable x instead of / in the
coefficient expression. Our intention is to let T —* oo, in which case/r (0 * f{t). —
Since the limiting process requires that a> = 2n/T — 0, for emphasis we replace
t ^ I
fT(x)e- Jlx
-' )n/uo
dx Aco (17.33)
- ,)
g(co, = 2^ I Mx)e- jalx dx (17.34)
fit) = lim
T->°° n=
2 gin^O), t)Aa) (17.35)
(Aeu-0)
464 Fourier Methods Chap. 17
But in the limit, fT —'/and 7— > oo in (17.34) so that what appears to be g(co, t) in
(17.36) is really its limit, which by (17.34) is
lim g(co, =
•*-££
t)
f( x ) e
-J^-» dx
)e-jMx-t) dx dm
fit)
lit
£[£ f(x (17.37)
W> = 5 x jo
f(x)e-'™ dx e " da> (17.38)
where we have changed the dummy variable from x to t. Then (17.38) becomes
f(/g>) = nnt)]
(17.41)
/(0 = ff- 1
PFC/Q))]
Also, (17.39) and (17.40) are collectively called the Fourier transform pair, the sym-
bolism for which is
The expression in (17.37), called the Fourier integral, is the analogy for a non-
periodic /(?) to the Fourier series for a periodic /(/). Equation (17.40) is, of course,
another form of (17.37). The Fourier transform corresponds, in this case, to the
Chap. 17 Fourier Methods 465
Fourier coefficients in the previous cases. Another description for these two analogies
is to say that the Fourier transform is a continuous representation (co being a con-
tinuous variable), whereas the Fourier scries is a discrete representation (/7co , for n an
integer, being a discrete variable).
As an example, let us find the transform of
fit) = e-°'u(t)
= P e- {a+M 'dt
Jo
or
$[e-°'u(t)] = \ . e
-e+j»)i
-(a + jco)
The upper limit is given by
since the expression in parentheses is bounded while the exponential goes to zero. Thus
we have
5[e-"'u(t)] = —+-^-
a jco
(17.43)
or
1
e-"'u(t)
a +jco
As another example, let us find the transform of the single rectangular pulse
T(jco) = j"~
f(t)e-^'dt
rt/2
= Ae~'°" dt
J-S.2
__ 2A /e
icoi 2
— -
e iw62 \
- co \ —pr )
466 Fourier Methods Chap. 17
f(t)
2 2
or
EXERCISES
17.6.1 Find SFfc-"!'!], where a > 0. /1/w. 2a/(C0 2 + a2)
0, elsewhere
= 0, |co|> 1
^/w. (1/71) Sa (0
6 6
F(yco)
2
Z. + jco
-|-
JUJ J
3 + jo
JUJ
-f-
6
(2 + Jco)(3 + jco)
Another operation involving Fourier transforms that we shall find useful is that of
time differentiation. Suppose we wish to find the Fourier transform of the derivative of
a function /(/). By definition, if
Therefore we have
tiM^jcoFiJoo) (17.46)
That is. the transform of the derivative of/is found by simply multiplying the trans-
form off by jco. This result may be extended readily to the general case
^M<->iJcoy¥iJco) (17.47)
where // = 0, 1, 2 We are assuming that the derivatives involved exist and that
the interchange of operations of differentiation and integration is valid.
From the results represented by (17.47), we may show that Fourier transforms can
be used to extend the concept of network functions, considered in Chapter 14, to the
.
case of nonsinusoidal excitations. For example, taking the Fourier transform of both
sides of (14.22) and making use of (17.47) and linearity, we have
[a n (j(o)
n
+ a^iijco)"-
1
+ . . . + a Jco + a ]\ o (jco)
-
--
[bjLjcoy I
b m ^Uco) m 1
+ . . . + bjco + boWMco)
The functions V and V, are the Fourier transforms of the output and input functions,
v and v From this result we may write
f
.
^>> - HUto)
V,Oco)
= bjjcoy + b m ^(j(or i
+ . . . + b
(17.48)
a n (jco)" + a n ^(jco) ni + . . +a
where, by (14.24), H(jco) is the network function of Chapter 14, evaluated at s =ja>.
This development indicates that the network function, defined previously as the
ratio of the output phasor to the input phasor, is precisely the same as the ratio of the
output transform to the input transform, and in the latter case the functions involved
do not have to be sinusoids. This generalization may be used also to find the time-
domain responses. For example, suppose the input v {t) is given by t
v t (t) = e-i<u(t)
^+ 2v = 6» t
or
yjijco) _ 6
H(yco)
\£]<o) jco + 2
3 + jco
the transform of the output is
V (jco) = HUcoyVtUco)
6
(2 + jco)(3 + jco)
By (17.45) this is
Chap. 17 Fourier Methods 469
v -(
»^(r^)- 6
(rns)
so that by linearity, or by (17.44). we have
Since the network function concept implies that there is only one input, the time-
domain responses obtained in this manner are responses of initially relaxed circuits.
(No initial energy is stored.) This is, of course, a disadvantage of the Fourier transform
technique.
We could develop general methods for using Fourier transforms to find time-
domain solutions, but there are many common functions which do not have Fourier
transforms, and the methods, as we have just noted, are limited to finding responses of
initially relaxed circuits. We prefer, therefore, to consider a more general transform
method, involving Laplace transforms, in the next chapter. As we shall see, Laplace
transforms are more versatile than Fourier transforms, in the sense that more func-
tions possess Laplace transforms. In addition. Laplace transforms may be used to
obtain the complete response, taking into account initial conditions.
EXERCISES
17.7.1 If x(t) is the input and y(t) is the output, find the network function, using
Fourier transforms, where
x = e' 2 'u{t)
PROBLEMS
17.1 Find the Fourier trigonometric series for the functions
(a) fit) = 1, < t < a
= 2,n<t<2n
/(/ + 2a) =/(/).
(b) fit) = e', —a < <n t
/(r + 4)=/(0
(c) /(r) = < <
/, r 1
/(/+ 1) =/(/).
17.3 Obtain the exponential series of the full-wave rectified sinusoid of Ex. 17.2.3
from its trigonometric series.
17.4 Obtain the trigonometric series of fit) given in Ex. 17.3.1 from its exponential
series.
.1
Ans. -=- + ^
—n2 fi
2j tt
—-
(-1)"
£—
+1
tx cos
(2/i -=—1)71/
-—
2 B (2a; 1 2
17.5 Find the first three terms of the forced response /(/) if v g it) is the function of
Prob. 17.4.
PROBLEM 17.5
17.6 Find the forced response i(t) in Prob. 17.5 if v g it) is the function of Prob.
17.1 (c).
17.7 Find the discrete amplitude and phase spectra of the functions of Prob. 17.1.
17.8 If fit) in (17.22) is the voltage across, or the current through, a 1-Q resistor,
then the instantaneous power is/ 2 (/), and the average power is
Chap. 17 Fourier Methods 471
P(t)dt= £ |f n P
Jo
I
e Jnwot e Jm<„o' dt = 0, 171, W = 0, ±1, ±2, . . .
Jo
unless m = — «.]
17.9 From Prob. 17.8 the quantity |c„| 2 is
the average power associated with the
frequency na> The plot of |c„| 2 versus frequency is a line spectrum known as
.
the discrete power spectrum of f(t). Plot this function in the cases in Prob.
17.2.
(a) /(/) = 1
a-
17.13 If the transform of Prob. 17.10 (a) is a network function, find the amplitude and
phase responses.
17.14 If the output is /(/) and the input is v g (t) in the circuit of Prob. 17.5, find the
network function.
17.15 If in the circuit of Prob. 17.5 the output is v c (t) and the input is v e it), find (a)
the network function H(jCO) and (b) v c (t) if r c (0)= and v g it) = e~'uit) V.
[Suggestion : Note that
472 Fourier Methods Chap. 17
(A- + l)(.v + 2) x + l x + 2]
17.16 Find the amplitude and phase response of a network having input x(t) and out-
put vC) related by
y" + 5/ + 4y = 2x'
17.17 If /(/) is the voltage across, or the current through, a l-Q. resistor, then f 2 (t)
is the instantaneous power delivered to the resistor. Show that if /(f) is given
by (17.40), then the total energy delivered to the resistor is
w= f p (r) dt = JL f I
F(jco) fdco
sometimes called the energy density. The energy in an output/(/) in the band of
frequencies co = a to co = b is then
1
"*
Find the total energy (a = —oo,b = oo) in the output of an ideal low-pass
having input e~'u(t) and amplitude response
filter
= 0, elsewhere
17.20 Find the network function if the output is y(t), the input is x(t), and they are
related by other variables q it q z> and q 3 by
dqi
-jf
= -« - 9. + i
y =Qz
18
LAPLACE TRANSFORMS
As we have seen in the previous chapter, the Fourier transform may be used to
extend the concept of network functions to cases where the excitation is nonsinusoidal.
It may also be used in certain applications to find the forced response to an input
function which possesses a Fourier transform. With the addition of the Fourier
transform to our arsenal we are able to add networks with certain nonsinusoidal,
nonperiodic excitations to our set of solvable circuits. We have thus made considerable
progress since we first solved resistive circuits with dc excitations in Chapter 2. Along
the way we have developed differential equation techniques for circuits containing
storage elements, phasor techniques for sinusoidal excitations, generalized phasors for
damped and exponential Fourier series methods for periodic
sinusoids, trigonometric
excitations, and now Fourier transforms for certain nonperiodic excitations.
In this last chapter it remains for us to consider a most elegant procedure, the
Laplace transform method, which can be applied to a much wider variety of excitation
functions than the Fourier transform method and which gives at once both the natural
and forced responses for a given set of initial conditions. In the process we shall define
a special and highly useful function, known as the impulse function, and use its
properties to obtain general time-domain solutions to any linear circuit whose excita-
tion possesses a Laplace transform.
The Laplace transform may be obtained as a generalization of the Fourier trans-
form and is developed in this way by many authors. However, to make this chapter
473
474 Laplace Transforms Chap. 18
18.1 DEFINITION
The Laplace transform, named for the French astronomer and mathematician Pierre
Simon, Marquis de Laplace (1749-1827), is defined by
The function F(s) is the Laplace transform of /(f) and is a function of the generalized
frequency, s + = a jco, We might note that our
considered earlier in Chapter 14.
tinguished from the two-sided or bilateral transform. The latter is defined by (18.1)
with the lower limit replaced by — oo. However, we shall not have need for this
generalization since for the circuits we shall consider the function /(f) is of interest only
for t > 0. In general, if/(0 has an infinite discontinuity at t — 0, we shall understand
that the lower limit is 0".
Comparing (18.1) with the definition of the Fourier transform of the previous
chapter, we see that
ne-"'f(t)u(t)] = £[/(/)]
The presence of the factors e~ at and u{t) accounts for the greater versatility of the
Laplace transform. For example, the Fourier transform of a function f(t) may not
exist because the condition
(18.2)
J" |/(0I<*<°°
fails to hold. It is possible, however, that for the same function /(r) a value of a may
be found so that
As a general rule, if/(/) = Ofor t < and ^[/(r)] exists, then the factor u{t) in (18.3) is
redundant, and the factor e~ at is not necessary for the existence of the integral. In this
case, the Laplace transform with s = jco is the Fourier transform.
As mentioned earlier, the Laplace transform can be used to take into account the
initial conditions prevailing in the circuit. To gain some insight into how this is
accomplished, let us consider the transform of the derivative f'(t), given by (18.1) as
£[f'0)]= rf'(t)e-"dt
Jo
[f /(f) and s are such that the integrated part vanishes at the upper limit, as will be the
case for virtually all functions we shall encounter in circuit theory, then in view of
(1 8. 1), this result becomes
-sF(s)-f(0) (1 8.4)
(it
[We are considering the case/(r) to be a continuous function on < <t oo.] Thus the
condition /(0) is automatically built into the transform of the derivative. This
initial
was not the case for the Fourier transform, as seen in (17.46).
To illustrate the computation of a Laplace transform, let us consider
/(/) = e -"u{t)
¥(s) = F e-"'e-"dt
Jo
1 -f(+,
If Re s — o > —a, then the value at the upper limit is zero, and we have
£[e°'u(t)] = 1
(18.5)
s + a
In general, we shall define /(r) in (18.1) as the inverse Laplace transform of F(.v),
e-"'u{t) (18.7)
s + a_
As another, simpler, but highly useful example, the transform of the step function
u{t) is given by
£" uit)
476 Laplace Transforms Chap. 18
This function further illustrates the versatility of the Laplace transform. The step
function does not have a Fourier transform since for it (18.2) does not hold. (That is,
EXERCISES
18.1.1 Find the Laplace transform of (a) sin ktu(t), (b) cos kt u(t), and (c)
(1 + 3e- 2 ')u(t)-
As was the case for the Fourier transform, the Laplace transform is an integral and is
where F and F 2 are the Laplace transforms of/, and/2 and c and c 2 are constants.
:
,
Thus we may obtain a variety of transforms from two known ones. For example, the
function of Ex. 18.1.1(c) may be easily transformed by writing
£[u(t) + 3e-
z
'u(t)] = £[u(t)] + 3£[e' 2 'u(t)]
= ^-
s
+ s^-
'
+2
As another example, since we have
and
£[e- Jk 'u(r)]
s -!- jk
-j- £[e
ik
'u(t)] + -j- £[e- Jk 'u(t)]
_s — TT.+
jk s + jk.
s
(18.10)
s
2
+ kl
£>[e-"'f(t)] = P e-"f(t)e-" dt
u+a)
--
I' f(t)e- ' dt
Jo
where F is the transform of/ As a consequence of ( 18.10), (18.1 1), and (18.12), we may
write the useful transforms
s a
£[e-"' cos bt u(t)} = -\-
(18.13)
and
b
£[e' a sin
'
bt u(t)] (18.14)
(s + a)
2
+ b
2
F(( 5 )} = ^
s
2
+ 2s + 5
65
F(s)
(s \y--2 2
1
F(s)
6Q -\- l)-6
(s + \y + 1
cV s+ \
l(s+ l)
2
+ 22 . is + l)
2
+ 22
-- [° fit -T)e J
'
dt
f(t-T)u(t-t)
(b)
FIGURE 18.1 (a) Function f(t) and (b) f(t) made zero for t < and
translated x units to the right
= e~" f(x)e~
iX
dx
Jo
or
£[/(/ - r)u(t - t)] = e-"F(s) (18.16)
y(t) =£->[G(*)F(j)]
G(s) \ gix)e~"dx
Jo
so that
Interchanging the order of integration and noting that u(t — x) = for x > t, we
have
G(s)FU)-- fV"|Ts(T)/(f-T)<fr dt
Jo L^o
G(j)F(5) = 'g(T)/(/-T)A
£[Jo
Thus the function r(r) is given by
This result is known as the convolution theorem, and the integral involved is the
convolution of g and/, symbolized by
We shall find the convolution theorem very useful later in the chapter.
EXERCISES
18.2.1 Show that the inverse transform is also a linear operation. That is,
£" I
[c 1 F 1 (5) + c 2 F 2 (s)] = c 1 £-i[F 1 (j)] + c 2 £"'[F 2 (*)]
480 Laplace Transforms Chap. 18
Therefore the last step leading to (18.15) is justified. [Suggestion: Invert both
sides of (18.9).]
18.2.2 Find the Laplace transform of (a) te~ 2 'u(t), (b) u(t — 3), and (c) f(t) = 1,
f(t)*g{t) = g(t)*f(t)
It is evident from the two sections of this chapter that we may compile a lengthy
first
fU) F(s)
1. S(t) 1
2. u(t)
3. e-"'iiU)
s + a
k
4. sin ktu(t)
s- + k2
s
5. cos kt u(t)
s2 + k2
h
6. e~ at sin bt ti(t)
(s + a) 2 + b2
8. MO h
common functions /(/) and their transforms F(^) which we are likely to encounter in
circuit theory. We have derived all of these previously except the first entry, a function
designated as 8(t) with transform F(s) = 1, which we shall define and consider in
Sec. 18.4.
To illustrate the use of the table, suppose we are required to find the inverse of the
Chap. 18 Lap/ace Transforms 481
transform
FO) -\
2
s 4 '
s I
9 s
Since, as discussed in Ex. 1 8.2. 1, the inverse transform is a linear operation, we have
+ +c-
l
_s
2
+y
By entries 3, 4. and 2 of the table we have
20 I- 10)
FO) (18.19)
+ 1)0 + 4)
There is no direct entry in Table 18.1 which we can use to obtain f{t) in this case;
however, we may obtain a partial fraction expansion of F(s) and apply the table to each
term in the expansion. Obtaining a partial fraction expansion is the opposite operation
of getting a common denominator. That is, we ask ourselves what simple fractions add
together to yield F(s). Since in this case the transform is a proper fraction (the numera-
tor is of lower degree than the denominator), the partial fractions will be proper
fractions and must therefore be of the form
20 + 10) s
FO)
+ 1)0 + 4) 5+1 +
The constants A and B are determined so as to make the second and third members
identities in s.
c+ 4) FW _ 2^±m _ db±p + B
Since these must be identities for all s, let us evaluate the first at s = — 1, which
eliminates B, and the second at s = —4, which eliminates A. The results are
2(9)
,4 = 0+ 1)F0)
3
= 6
2(6) -4
B = + 4)F(s)
-3 =
482 Laplace Transforms Chap. 18
Therefore we have
F(.)
s + I s + 4
and, by Table 18. 1,
9s 3
F(s) =
(s + l)(s
2
+ 2s+ 10)
We note that the degree of the numerator is the same as that of the denominator, so
that long division is required. It is not necessary to actually perform the division since
it is evident that it will result in 9 plus a proper fraction. Therefore let us write
9c 3
F(*) =
(s + \)(s
2
+ 2s+ 10)
= 9-1 A
+ s+ +I
Bs + C (18 20)}
1 s
2
+ 2s + 10
K
'
We know that the linear denominator term has a constant numerator and that the
quadratic denominator term has, in the most general proper fraction case, a linear
numerator, as indicated. We could factor the quadratic into two linear terms and
proceed as in the example of (18.19). However, this will result in complex coefficients,
so we choose to leave the quadratic factor intact.
We may find A, as in the previous example, from
~9
A = (s + l)F(s)
"1-2+10 -1
,«-,
9s 3 = 9(5 + l)(s
2
+ 2s + 10) - (s
2
+ 2s -f 10) + (Bs + C)(s + 1)
(We have substituted in the value of A.) Since this is an identity in s, the coefficients of
like powers must be equal in both members of the equation. Thus the coefficients of
Chap. 18 Laplace Transforms 483
Z
5 arc
0=18 + 9-1 + 5
from which fi = — 26. The coefficients of 5° are
= 90 - 10+ C
yielding C — —80. Therefore we have
F( _ 1 __26, + 80
5
s + 1 5
s
2
+ 2s + 10
which we rearrange as
F(s) = 9 26,,
s +]
1 L(J + 1) + 3
2
_ (*+ I)
2
4
F(5) =
(5 + 1) (5
2
+ 2)
then the procedure is slightly more complicated. Evidently in this case the partial
fraction expansion is
B
F(5)
(s + 1) (5
2
+ 2) (s + l)
2 +s+
'
1 '
s + 2
(18.21)
since the terms on the right combine to yield the proper common denominator. There
is no need to use a linear numerator for the second-degree denominator term, since
K,s +K _ 2 K,(s + 1) +K -K 2 x
(s + l) 2 "
(s + l) 2
_ A", K —K 2 x
5+1 '
(S + l)
2
/* = (* + l)
2
F(s)| =4
U=— i
C= (s + 2)F(j) = 4
1
Equating coefficients of s yields
5 + 4 =
or B = — 4. The transform is then
Hs)
(s +
4
l)
2
s
4,4+
+ 1 '
s 2
The first term may be inverted by means of entry 8 in Table 18.1 and (18.12) and the
other terms by entry 3. The result is
As observed earlier, an explicit relation for the inverse transform is given in Prob.
18.6, but for our purposes we may generally use partial fractions and Table 18.1 to
obtain the inverse transforms.
EXERCISE
18.3.1 Find the inverse transform of
(a)
(s !- \){s + 2)(s + 3)
s nnA
(b)
(s 2 + \){s 2 + 4)'
(c)
s* + 5s + 21 s 2 + Als +
3
78
{s 2 + 9)(s 2 +2^ + 5)
known in the literature as the impulse function. It is not a function in the conventional
sense, but it can be defined and
its properties established by using generalized, or
distribution, function theory. In any case, it is a very useful function in circuit theory,
and we shall give a plausibility argument for its existence in this section.
Let us begin by considering the finite pulse of Fig. 8.2, centered about the origin, 1
/W =i-> -f«<i
= 0, elsewhere
Chap. 18 Laplace Transforms 485
f(t)
/a
-a/2 a/2 t
The area under the pulse is 1 and remains fixed regardless of the value of a. If a is
made smaller, then the base of the pulse shrinks and its height increases, maintaining
the constant area of 1. toward zero, the pulse approaches an
In the limit as a tends
infinite pulse occurring over zero time but still associated with an area of 1. Such a
function is called an impulse function, or a unit impulse function, to indicate its rela-
tionship to the unit area, and is symbolized by 5{t). Graphically it is represented by an
arrow erected at t — 0, as shown in Fig. 18.3(a). A more general function is 8{t — t),
6(t)
(a) (b
5(t) - 0, I ^
(18.22)
8(t) dt = 1
J
or more generally by
8(t — t) = 0, t^x
(18.23)
d\t - t) dt = 1
J
Thus the impulse is zero everywhere except at its point of discontinuity, at which it
has an area of 1 concentrated. Physically, the impulse describes very well a force of
very large magnitude exerted for an extremely short time and in spite of its abstract
mathematical definition is quite a useful function.
An important property associated with the impulse function is the sampling
486 Laplace Transforms Chap. 18
property, described by
where a < x < b and/(?) is continuous att = r. This property may be made plausible
by noting that since 8{t — x) is zero except at t — x, we may write
f
Ja
f(t)d(t-z)dt=- r
Jr — e
f
-x)dt
f(t)d(t-x)
b t + f
\
f(t)d(t -x)dt = f(x) \
8(t - x) dt
Ja Jx-€
= m
The sampling property is very useful in determining the Laplace transform of the
impulse function, given by
(Recall that in the event of an infinite discontinuity of the integrand at 0, we are using
-
as the lower limit.) By the sampling property, with x = 0, we have
f(t)d(t) = f{0)S(t)
where /(/) is continuous at / = 0. This is plausible since 5(t) = for / ^ 0. It arises,
for example, when we differentiate functions that contain u(t) as a factor, such as in
Ex. 18.4.3.
EXERCISES
18.4.1 Evaluate the integral
)
(/
2
+ 3 cos 2t)5{t) dt
Ja
a
so that
.. ... . chid)
,im /(')
-0
= —JT
«'
a
<5(/) =
^ )
= f'(t)u{t) + /"((>)<$(/)
18.4.4 Use the result of Ex. 18.4.3 and (18.4), which, because of the discontinuity, we
write as
£[/'(/)] = *£[/(/)] -/(0-)
The Laplace transform, like the Fourier transform, may be used to solve differential
equations. However, the Laplace transform method has the advantage of yielding the
complete solution with the initial conditions accounted for automatically, as we shall
see in this section.
In Sec. 18.1 we have already established the result of (18.4), which we repeat here
as
£.[f"(t)] = s£[f'(t)]-f'(0)
or. by (18.26).
£[/"(')] - 2
s ¥(s) - sf(0) - /'(0) ( 1 8.27)
We may replace /by/' again in (18.27) to obtain £[/'"(()], and so forth, with the
general result being
£[/
(fl,
(0] = s"F(s) - s" </(0) - 5"" 2/'(0) -- • • • -/"-"(0) (18.28)
where (n)
The functions//', ,/'" n are assumed to be
/ is the /?th derivative. . . .
488 Laplace Transforms Chap. 18
(n)
continuous on (0, oo), and f is continuous except possibly for a finite number of
finite discontinuities. As in Ex. 18.4.4, if there is an infinite discontinuity present in
.y(0) = 1 , a-'(O) = 2
s
2
X(s) - s - 2 + 4[sX(s) - 1] + 3X(j) = 1
from which
2
+ 8s + 13
x( W_ v
(s +
s
V)(s + 2)(s + 3)
Y 3 1 1
Ms)
, ,
s + 1 s + 2 s + 3
(Since we are interested in / > 0, we have not needed to consider the step function
involved in the transforms.)
We may also handle certain integrodifferential equations directly without differ-
entiating to remove the integrals. Toward that end we need the transform of g(t),
given by
Differentiating, we have
g\t)=fit) (18.30)
£[g'(t)} = s£[g(')}
or
£[*(')] =- y £[g'(0]
Chap. 18 Laplace Transforms 489
FC£)
£ t)dt (18.31)
s
Jo
To illustrate the use of (18.31), suppose we have an RLC series circuit with
R = 2Q, L = 1 H, and C= | F, excited by a unit step function, v g (t) = u{t) V.
Then the current i{t) satisfies the equations
|- + 2H 5 f'/<ft = tt(0
1(0) =
Transforming, by (18.26) and (18.31), we obtain
«*) = =r
2
+ + 2
+
s 2.y 5 "
2L(^+ l) 4.
Suppose in the general case that the input v t and output v are related by the
differential equation
d"v,
'dr [
d"'v
dt
-• +«i^ + flo«.
-{-b^ + boV;
fife
and that the initial conditions are all zero; that is,
t'o(0)
rfy.(0) tf-'^O) __ v{0) _ dvjjO)
dt dt"' 1
^ dt"'-'
5"-
(a„s" + fl B - 1
1
- •
«,* -: fl )V„(j)
HUj - - U8 JZ;
a n . lS - +a + '
V,(5) a n s" + 1
+ lS aQ
490 Laplace Transforms Chap. 18
Comparing this result with (17.48), obtained in the previous chapter, we see that H(s)
is the network function, with/a> generalized to s. Or, going back to (14.24) of Chapter
14. H(.v), the ratio of the Laplace transforms of the output to the input, is precisely the
network function as it was Moreover, since admittances and
originally defined.
impedances are network functions, they are precisely the same in the Laplace trans-
form domain as they were in the case of the generalized phasors. With Laplace
transforms, however, the situation is more general. We may have any function that is
Laplace transformable as the input or output, instead of being restricted to damped
sinusoids.
To illustrate with the last example, if the input of the RLC series circuit is v g (t)
HW-^--
V g (s)
-J-
Z(s)
or
H(j) = I
sL +R+ (\/sC)
_ 5
~ (18.33)
s
1
+ 2s + 5
Then
\ (s) = H{s)
In this case v (t) is called the impulse response (response to an impulse) and is denoted
by h(t). That is,
v {t) = h(t) = £~ 1
[H(s)} (18.35)
The impulse response is thus very important since its transform is the network func-
tion itself.
Another point should be made in this section. Since, in general, \ (s) is by (18.34)
a product of two transforms, we may write immediately the output v (t) by means of
the convolution theorem of (18.17). The result is
Therefore if we know the network function and the input, we may find the forced
response directly.
As an example, for the RLC series circuit considered earlier in this section, the
i(/) = A(0*"(0
I e
r
(cos 2t — \ sin 2x)u(r)u(t — x) dx
•'o
I e
T
(cos 2t — £ sin 2t) f/r
Jo
= £e~' sin It
as before.
The impulse response of this circuit is. of course, /?(/). As in the case of the step
response, with an impulse input there is no initial stored energy at / = 0~. Thus /;(/)
2x' + / - y= \
x' +x+y= o
x(0) = 2, r(0) = -
Transforming, we have
v, v 2^ + 2
y(t) = —1 — 4 sin /
492 Laplace Transforms Chap. 18
EXERCISES
18.5.1 Solve for .v(r), for / > 0, by using Laplace transforms:
(a) x" + 3.v' + 2.y = <5(/).
(b) x" + Ax' + 3.v = 4e~'
.v(0) = ,v'(0) = 4. Ans. (a) e~' - e~ l >, (b) e~'{2t + 7) - 3e~ 3 '
H(.) -
2( * + 2)
{S + 1)(5 + 3)
find (a) //(/), (b) the step response, and (c) the forced response to an input e~ 2 '.
Ans. (a) (<?"' + e~ 3,
)u(t), (b) j(4 - 3e~' - <T ')«(0,
3
When dealing with phasors we were able to omit the steps of writing the differential
equations by going directly to the phasor circuit. The circuit was analyzed by resistive
circuit techniques, the phasor output was found, and from this the time-domain
solution was obtained. We may perform a similar type of analysis with Laplace
transforms by first obtaining a transformed circuit. The rest of the procedure is similar
to that of the phasor method except that with Laplace transforms we are able to deal
with more general functions and obtain the complete solution with the initial condi-
tions satisfied as well.
To see how this may be done, let us consider first a resistance R, with current i R and
voltage v R for which
,
vR = Ri R
which is represented by the transformed resistor element of Fig. 18.4(a). Next, let us
consider an inductor L for which
vL = A dt
Transforming, we have
VL(s) = sLIL (s) - LiL (0) (18.38)
This result is represented by the transformed inductor element of Fig. 18.4(b). The
initial condition is taken into account by the included voltage source. In the case of a
Chap. 18 Laplace Transforms 493
I (s)
R
ys)
— 6
capacitor C we have
i c dt '
r r (0)
Transforming, we obtain
since for t > the transform of a constant A' is K/s. The last result is represented by
the transformed capacitor of Fig. 18.4(c). The initial condition is provided for by the
included voltage source.
If we have dependent sources, we may transform them in the same manner. For
example, if we have a controlled voltage source defined by
f, = Kv 2
V,(.v) = K\ 2 (s)
which in the transformed circuit is a source controlled by a transformed variable.
We may write the time-domain equations of a circuit and solve them by Laplace
transform methods, or we may replace the circuit by the transformed circuit. In this
case the passive elements are replaced by transformed elements such as those of Fig.
18.4. the controlled sources are appropriately transformed, and the independent
sources are labeled with the transforms of their time-domain values. Then we may
solve the transformed circuit using resistive circuit methods.
As an example, suppose we require /'(/), for / > 0, in Fig. 18.5(a), given that
;(0)= 4A and r(0) = 8 V. The transformed network is shown in Fig. 18.5(b), from
which we have
+ 3)] + -
W_
Ir
[2/(s
3 +s+
4
(2/s)
(8/5)
(18.40)
i(t)
3n I H
2e" 3t V
© v(t)
Ks)
3
-nm^ —Qy
hO s
di
+ 3/ i di + 8 = 2e~
dt
/(0) = 4
Transforming, we have
given by (18.40).
+ 31(5) + j-I(s) + A=
^
Jf we are interested in nodal analysis, we need to solve (18.37), (18.38), and (18.39)
explicitly for the currents. The results are
I*(*)
R
As the reader may verify, the transformed elements describing these equations are
shown in Fig. 18.6. The labels on the passive elements are impedances, as in Fig. 18.4.
Chap. 18 Laplace Transforms 495
+
T IIrIs) I (s) I (s)
L c
VR (s) \ R )!j
r r v (s)
Cp Cvc (0)
cii c
All the network theorems that apply to resistive and phasor circuits apply to the
transformed circuit. An added advantage in the latter case is that initial conditions may
be incorporated. As an example, the reader may verify that the elements of Fig. 18.6
are the Norton equivalents of those of Fig. 18.4.
EXERCISES
18.6.1 Solve Prob. 14.19 using the transformed circuit method if r<0) = 10 V.
Ans. 10(2 - e~') + {e~' sin It V
18.6.2 Find the voltage <•(/), for / > 0, across the terminals of the RLC parallel circuit,
excited by a current source i
g (t)
= e' 1 A,
'
using the transformed circuit
method. The passive elements are R = 4fi, L = 5 H, and C == ^o f the ^
PROBLEMS
18.1 Find the Laplace transform of the functions
(a) «(/) - //(/ - 1).
(b)
(s + \\s + 2)
(c)
s + 1
(d)
(s + \)(s + 2)
[Suggestion : Write the transform as an infinite series of integrals over (0, T),
(T, 2T), . . . , and sum.]
18.4 Use the results of Prob. 18.3 to obtain the transforms of the functions
(a) /(f) = 1, <t< 1
= 0, < < 1 f 2
fit + 2) = /(f).
(b) /(*) = sin |
/|.
18.5 Note that if />(/) is a pulse of some shape of finite duration, occurring on
< < t T[p(t) = elsewhere], and that if/(/) is defined by
£[p(t)]
£[/(/)] =
1 - e-
(I — e ~s\l
s(\ - e' 2s )
18.6 In the Fourier integral (17.37), replace /(f) by e~"'f(t), and show that if/(f) = 0,
for t < 0, then
fit) - ^7 e "F(s) ds
This is the inverse Laplace transform analogous to (17.40) in the Fourier trans-
form case and is valid if a is sufficiently large to ensure that (18.3) holds.
18.8 Solve Prob. 18.7 by finding F(s) and G(V) and using the convolution theorem.
Chap. 18 Laplace Transforms 497
His)!
hit) =£-'[H(.v)]. r(0 = £-'|^]
so that, by (18.31),
2a-' + 4a- +/ + 7 y =-
x x +y + ly = d(t)
498 Laplace Transforms Chap. 18
18.16 Find the network function and the impulse response if the output is /'(/).
ilt)
PROBLEM 18.16
His)
s + 1
and no initial stored energy to the input /(f) of Prob. 18.4(a). Suggestion:
Find the inverse transform involved by the procedure of Prob. 18.5, but note
that if p(t) ^ 0, for t > T, then the inverse is not periodic but is
18.20 Find the network function of Prob. 9.22, and show that the circuit is a band-
passfilter. Find also (o and Q.
18.21 With the presence of impulse currents or voltages in a circuit it is possible,
theoretically, to change inductor currents and capacitor
instantaneously
voltages. Demonstrate this by finding v(i) and /'(/) for — oo < / < oo in
the given circuit. The switch is closed at / = and v(0~) = 0.
i(t)
6V
PROBLEM 18.21
Find the network function and the amplitude response, and thus show that
the circuit is a second-order Butterworth low-pass filter with co c = 1 rad/s.
APPENDIX
A
DETERMINANTS AND
CRAMER'S RULE
(A.l)
a,,
'21 a,.
'22
In this case the determinant is a 2 x 2 array, with two rows and two columns and a
value A defined to be
As an example, consider
1 2
-3 4
which is given by
A = (1)(4) - (2)(-3) = 10
499
500 Determinants and Cramer's Rule App. A
12 0.3
22 «23 (A.4)
32 #33
has three rows and three columns. It may also be evaluated by a diagonal rule, given by
aX 13
^a><A
= («llO22033 + 01202303! + 01 3 03202l)
Thus the value of the determinant is a difference of products of elements down the
diagonals to the right and products of elements down the diagonals to the left.
1 1 1
2 -1 1
-1 1 9
= -7 (A.6)
1 1
A 2l
1 2
In other words, the cofactor is the signed minor, the minor multiplied by ± 1 with the
App. A Determinants and Cramer's Rule 501
sign depending on whether the sum of the row number and column number is even
or odd.
The value of a determinant is the sum of products of the elements in any row or
column and their cofactors. For example, let us expand the determinant of (A. 4) by
cofactors of the first row. The result is
«nCii + a xl C v ,
A = fl I1 (-l)' + M +a ll 12 (-iy+*A 12
A = au — a. + «u
a-n a,
A = a n {a22 a 33 - a23 a 32 ) - a 12 (a 21 a 33 - a 2i a n ) + a 13 (a 21 a 32 - a 22 a 31 )
2 - +
H-l) 1
-1
H-D 2 :
2(— 1) 3+;
(1 + + 2(- -7
x, - 2.v = 2 5
(A.8)
6x, + x = 2 4
We define the determinant of the system as the determinant A whose first column is the
coefficients of x u whose second column is the coefficients of .y 2 , etc. In the case of the
system of (A.8) we have
1 -2
13 (A.9)
6 1
502 Determinants and Cramer's Rule App. A
A
(A. 10)
x, =
etc.
example of (A. 8), the right members of the equations are 5 and 4 so that
5 -2
A, = 13
4 1
1 5
A, = -26
6 4
x, = A 1 _13_ 1
13
-26
2
= -2
2
"A 13
x t
+ x2 + x3 = 6
2xj — x2 + .v 3 = 3 (A.ll)
x2 + 2x 3
6 1
3 -1 1
7 1 2 -7
1 1 1 -7
2 -1 1
-1 1 2
1 6 1
2 3 1
x, = -1 7 2 -14
-7
App. A Determinants and Cramer's Rule 503
and
I 6
1 3
7 -21
-7 =
I
3
-7
APPENDIX
B
GAUSSIAN ELIMINATION
-.v, + x2 + 2x 3 = 7
equation by subtracting twice the first equation from the second and adding the
first equation to the third equation. We shall then have the first equation and the
two new equations neither of which contains .v,. The result is
.y, + x2 -\- x3 = 6
-3x 2 - x3 = -9 (B.2)
2x 2 + 3.y 3 = 13
.Y, + x2 + x3 = 6
x2 + jx 3
= 3 (B.3)
2.y 2 + 3.y 3 = 13
We next eliminate x 2 from all the equations except the first two. In the simple case
of (B.3), this means eliminating x 2 from the third equation. This may be done by
504
App. B Gaussian Elimination 505
•v, + x2 x3 =6
xm [
jX 3 — 3 (B.4)
7
J*3
.V, + a- 2 + x3 = 6
v, \x 3 = 3 (B.5)
x3 = 3
In a higher-order system we would repeat the procedure. That is, eliminate .v 3 from
all the equations except the first three. Eventually we reach the point, as in (B.5),
3 - j(3)
= 2 (B.6)
The known values of x2 and x 3 are then substituted into the first equation of (B.5),
yielding
•v, = 6 — x2 — x3
=6-2-3
= 1 (B.7)
In this example, this last step completes the process, since all the unknowns have been
found.
The steps in the Gaussian procedure may be put in compact form by leaving out
the symbols for the unknowns, the addition operation signs, and the equal signs. That
is, (B.l) may be represented by
2-113 (B.8)
1 12 7
Such an entity is sometimes called a matrix. It has rows and columns like a deter-
minant, but it is is 3 x 4) and has no
not necessarily a square array (the example given
number associated with it. The matrix merely represents the equations. That is, its
first row means one .v, plus one x 2 plus one x 3 equals six, etc.
506 Gaussian Elimination App. B
1 1 1 6
-3 -1 -9 (B.9)
2 3 13
1 1 1 6
1
i 3 (B.10)
2 3 13
This represents (B.3).
Subtracting twice the second row of (B.l 0) from the third row gives
1116
(B.ll)
.0 $ 7.
1116
1
\ 3 (B.12)
13
The last result is equivalent to (B.5) and may be used, as in (B.6) and (B.7), to obtain
the solution.
Alternatively, we may continue the elimination process as follows. In (B.12),
subtract from the second row ^ times the last row, and subtract the last row from the
first row. This results in
110 3
10 2
13
Finally, subtract the second row from the first, obtaining
1 1
1 2 (B.13)
1 3_
C
COMPLEX NUMBERS
From our earliest training in arithmetic we have dealt with real numbers, such as 3,
— 5, $, 7i, etc.. which may be used to measure distances in one direction or another
from a fixed point. A number such as x that satisfies
x1 = -4 (C.l)
is not a real number and is customarily, and ijnfnrtnnatply, rai led an imaginary
number. To deal with imaginary numbers, an imag inary unit, denoted by /', is defined
j = J~\ (C2)
Thus we have j 2 = — \,j = —j,j4 = 1, etc. (We might note that mathematicians use
3
the symbol i for the imaginary unit, but in electrical engineering this might be confused
with current.) An imaginary number is defined as the product of/ with a real number,
such as x = /2. In this case x 1 = (y'2) 2 = — 4, and thus .y is the solution of (C. 1 ).
A complex number is the sum of a real number and an imaginary number, such as
A = a + jh (C3)
where a an d b are real. The complex number A has a real part, a, and an imaginary
which are sometimes expressed as
part, d,
a = Re A
b = Im A
It is important to note that both parts are real, in spite of their names.
The complex number a - jb may be represented on a rectangular coordinate
plane, or a complex plane, by interpreting it as a point (a, b). That is, the horizontal
507
508 Complex Numbers App. C
coordinate is a and the vertical coordinate is b, as shown in Fig. C.l, for the case
4+y"3. Because of this analogy with points plotted on a rectangular coordinate
system, (C.3) is sometimes called the rectangular form of the complex number A.
IMAGINARY AXIS
3 • 4 + 3 ,
12 3 4
REAL AXIS
FIGURE C.1 Graphical representation of a complex number
A = a + j b
r = V« + b2 2
(C.4)
= tan-' —
and that
a = rcosO
(C.5)
b = r sin 9
App. C A Complex Numbers 509
(C.6)
is sometimes denoted b\
—
which is called the polar form. The number
r
/• is called the amplitude, or magni tude, and
.v = ang A - arg A
One may easily convert from rectangular to polar form, or vice versa, by means of
(C.4) and (C.5). For example, the number .4, shown in Fig. C.3, is given b>
A = 4 +./3 = 5/36.9
since by (C.4)
= J¥ + 3 = 5
r
2
Q = tan | = 36.9 1
A* = a - jb (C.7)
|
A* |
--=
Ja 1
+ (-b) 2 = Ja 2
+ b1 = \A\
and
arg A* - tan" 1
(—\ = — an -arg A
We may note from the definition that if A* is the conjugate of A, then A is the
conjugate of A*. That is, (A*)* = A.
The operations addition, subtraction, multiplication, and division apply to
complex numbers exactly as they do to real numbers. In the case of addition and
subtraction, we may write, in general,
This may also be done graphically, as shown in Fig. C.4(a), where the numbers .4 and
B are represented as vectors from the origin. The result is equivalent to completing the
parallelogram or to connecting the vectors A and B in head-to-tail manner, as shown
in Fig. C.4(b), as the reader may check by comparing the numbers. For this reason,
A + 8
(a) (b)
A =
+ jb = i\a cos 0, +./>i sin 0,
(C.ll)
B = C + jd = r 2 cos 9 2 + ji'i sin Qz
we have
AB = (a +jb)(c +jd) = ac + jad + jbc + j 2
bd
= {ac - bd) +j(ad+ be) (C12)
/,,-.[cos(0, ;
; ) i
./sin «9, + : )]
and hence we may multiply two numbers by multiplying their magnitudes and adding
their angles.
From this result we see that
AA* = (r/9)(r/-9) = r
2
/0=\A\ 2 /0
2 2
Since |
/J
|
/0 is the real number |
/I
|
, we have
|/*|
2
= /*/** (C.14)
A_
= a + jb
B ~
TV
c + jd
an irrational denominator, since /
results in ^J = '— l . We may rationalize the denomi-
nator and display the real and imaginary parts of N by writing
which is
^f = ^/^ (C6)
and
a + jb == c -f jd
then
a — c = j(d — b)
which requires
a = c, b — d
1 +x+j(S-2x) = 3+jy
then
1 +x = 3
8 - 2x = y
or x = 2, y = 4.
APPENDIX
D
EULER'S FORMULA
To derive Euler's formula, an important result, let us begin with the quantity
^ = V(cos0^ ;sin0)-./g
as may be seen from (D.I). The variables in the last equation may be separated,
yielding
g
Integrating, we have
lng = j»H K (D.2)
In 1 = = +K
or K— 0. Therefore we have
In g = j6
or
g = e'
e
(D.3)
513
—
514 Eu/er's Formula App. D
e
je
= cos0 + y sin 9 (D.4)
pie ~j6
i
p
cos 9 = - l_£_ (D.6)
Similarly, subtracting (D.5) from (D.4) and dividing the result by 2j, we have
pie
e _ p-ie
e
sin 9 = (D.l)
2
A = r[9 (D.8)
of the complex number
A=a+jb (D.9)
a = /• cos 9
(D.IO)
b = r sin 9
re'
6
= r(cos 9 + j sin 9)
= r cos 9 + j'f sin 9
which, by (D.IO), is
re'
s
= a+jb (D.I l)
A = r[l=re ie
(D.l 2)
This result enables us to easily obtain the multiplication and division rules, given
A = rJ9, = r e ie
x
'
App. D Euler's Formula 515
and
then
AB -(rJOMrz/Oz)
r l r z e ne,+6 ' )
-- r r 1 /B
l l
\9 1
^/g.-fla
8
specified by /cot or e iu". Its real part is the projection on the horizontal axis, given by
l
cos cot, and its imaginary part is the projection on the vertical axis, given by sin cot.
That is.
e'°" = cos cot + j sin cot
which is Euler's formula. The projections trace out the cosine and sine waves, as
shown, as the vector rotates.
Bandwidth, 393
Battery, 1
Ammeter, 40
Ampere, A.M., 5, 142
Ampere, 3, 5
Ampere turn, 436
Amplifier circuit, 54 Candela, 2
519
11
520 Index
inverse, 464
pairs, 464
Four-wire system, 340
Effective value, 306, 3 17 Frequency, 244
Electric field, 135 angular, 244
Electric filter (see Filter) complex, 360, 363, 365
Electromotive force, 7 cutoff, 392
Electron, 4 domain, 258
Element (see Circuit element) fundamental, 448
Energy, 7, 8 generalized, 365
in capacitor, 136 natural, 378
density, 472 neper, 361
in inductor, 145 radian, 244
in mutual inductors, 425 resonant, 244
storage element, 132 response, 389, 390
Envelope of spectrum, 462 scaling, 404
Equivalent circuit of transformer, 440 spectrum, 461
Euler, L.,210 continuous (see Amplitude response or Phase
Euler's formula, 210, 513, 514 response)
Even function, 452 discrete, 461
Excitation, 12 line, 461
Exponential current, 6 Full-wave rectified sinusoid, 456
Exponential Fourier series, 456, 457 Fundamental harmonic, 448
Fundamental period, 307
G
Farad, 133
Gain of a filter, 412, 414, 415
Faraday, M., 133, 142
Gaussian elimination, 66, 504
Feedback mode, 54
Gauss-Jordan method, 506
Fibonacci number, 131
Generalized node, 68
Filter, 284, 389, 392
Giga-, 3
bandpass, 392
Graph, 117, 118
band-reject, 394
connected, 118
gain of, 412,414, 415
Ground, 62
high-pass, 394
low-pass, 393
Flux linkage, 142
Force, 3 H
Forced response, 174, 212
Forcing function, 172 Half-range series, 455
complex, 251, 253 Half-wave rectified sinusoid, 455
damped sinusoidal, 360, 364 Harmonic, 448
sinusoidal, 243 Harmonic oscillator, 240
Fourier, J.-B. J., 446 Heaviside, O., 182
Fourier coefficients, 447 Helmholtz, H.S. F., 101
522 Index
linear, 142
element, 90
Linearity, 151, 153
practical, 150
Joule, J. P., 3
M
Joule, the, 3
Parallel connection, 30
N
Parallel resonance, 396
Parameters:
Nano-, 3
admittance, 382
Napier, J., 361
hybrid, 382, 384
Natural frequency, 207, 378
impedance, 381
Natural response, 160, 174, 205
transmission, 382, 388
from network function, 375
Parseval's theorem 47 1 , , 472
Negative phase sequence, 339
Partial fraction expansion, 481
Neper, 361
Passband, 392
frequency, 361
Passive element, 10
Network, 1
Period, 244
function, 360, 370, 371
Periodic extension, 455
one-port, 380
Periodic function, 244
theorems, 89, 288
effective value of, 317, 318
topology, 117
Laplace transform of, 496
two-port, 380
Per phase analysis, 343
Neutral terminal, 335, 338
Phase, 245
Neutral wire, 336
angle, 245
Newton, Sir I., 3
current, 341
Newton, the, 3
impedance, 340
Nodal analysis, 61 , 62, 280
response, 389, 390
Node, 19, 118
voltage, 338
generalized, 68
Phasor, 243,256
reference, 62
circuit, 269-71
voltage, 62
diagram, 294
Noninverting input terminal, 52
generalized, 363, 365
Nonlinear circuit, 92
reference, 295
Nonlinear resistor, 16
Physical resistor, 43
Norton, E. L., 101
Pico-, 3
Norton's theorem, 99, 102, 288, 290
Planar circuit, 73
Pliers entry, 378
Polar form of complex number, 250
O Pole, 373, 374
Pole-zero plot, 374, 401
Odd function, 452 Polyphase source, 333, 339
Ohm.G.S., 15 Port, 380
Ohm, the, 15 Positive-phase sequence, 339
Ohmmeter, 40 Potential coil, 327
Ohm's law, 15, 16 Potential difference (see Voltage)
1
524 Index
Power, 7, 8 Response, 12
apparent, 319 complete, 204, 218, 229
average, 306, 307 forced, 160, 174,204,212
complex, 306, 323 impulse, 490
factor, 306, 319 natural, 160, 174,204,205
of three-phase load, 342 Rms value, 306, 317
instantaneous, 8 Root-mean-square value, 317
measurement, 327, 353, 355 steady-state, 174
peak, 306 transient, 174
quadrature, 324
reactive, 324