0% found this document useful (0 votes)
203 views350 pages

Libro Zeng

This document is a tutorial on electric circuits written by Gengsheng Lawrence Zeng and Megan Zeng. It contains 27 chapters that cover fundamental concepts in electric circuits such as voltage, current, resistance, Kirchhoff's laws, circuit analysis methods, capacitors, inductors, and more. Each chapter provides clear explanations of the topics in a reader-friendly format with examples and exercises. The intended audience is students looking to better understand electric circuits.

Uploaded by

Nose Nose
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
203 views350 pages

Libro Zeng

This document is a tutorial on electric circuits written by Gengsheng Lawrence Zeng and Megan Zeng. It contains 27 chapters that cover fundamental concepts in electric circuits such as voltage, current, resistance, Kirchhoff's laws, circuit analysis methods, capacitors, inductors, and more. Each chapter provides clear explanations of the topics in a reader-friendly format with examples and exercises. The intended audience is students looking to better understand electric circuits.

Uploaded by

Nose Nose
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 350

Gengsheng Lawrence Zeng

Megan Zeng

Electric
Circuits
A Concise, Conceptual Tutorial
Electric Circuits
Gengsheng Lawrence Zeng • Megan Zeng

Electric Circuits
A Concise, Conceptual Tutorial
Gengsheng Lawrence Zeng Megan Zeng
Utah Valley University University of California, Berkeley
Orem, UT, USA Berkeley, CA, USA

ISBN 978-3-030-60514-8 ISBN 978-3-030-60515-5 (eBook)


https://doi.org/10.1007/978-3-030-60515-5

# The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland
AG 2021
This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the
material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilms or in any other physical way, and transmission or information
storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology
now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this
book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

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

Are you a student who is looking to supplement what you are learning in class? Or
are you simply interested in electric circuits? Electric Circuits: A Concise Concep-
tual Tutorial gives you an opportunity to understand fundamental electrical engi-
neering concepts. This book is written in a reader-friendly format like a pictorial
dictionary, and you can directly jump to any topic you want to learn more about
without having to read the entire book sequentially. We hope that this book will help
save your time in grasping difficult concepts in electric circuits.
Good luck and have fun!

Orem, UT Gengsheng Lawrence Zeng


Berkeley, CA Megan Zeng
2020

vii
Contents

1 Voltage, Current, and Resistance . . . . . . . . . . . . . . . . . . . . . . . . . . 1


Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2 DC Power Supply and Multimeters . . . . . . . . . . . . . . . . . . . . . . . . . 9
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3 Ohm’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
4 Kirchhoff’s Voltage Law (KVL) . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
5 Kirchhoff’s Current Law (KCL) . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
6 Resistors in Series and in Parallel . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
7 Voltage Divider and Current Divider . . . . . . . . . . . . . . . . . . . . . . . 43
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
8 Node-Voltage Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
9 Mesh-Current Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
10 Circuit Simulation Software . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
11 Superposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
12 Thévenin and Norton Equivalent Circuits . . . . . . . . . . . . . . . . . . . . 81
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
13 Maximum Power Transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

ix
x Contents

14 Operational Amplifiers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
15 Inductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
16 Capacitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
17 Analysis of a Circuit by Solving Differential Equations . . . . . . . . . . 119
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
18 First-Order Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
19 Sinusoidal Steady-State (Phasor) . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
Exercise Problems with Solutions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
20 Function Generators and Oscilloscopes . . . . . . . . . . . . . . . . . . . . . . 137
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
21 Mutual Inductance and Transformers . . . . . . . . . . . . . . . . . . . . . . . 149
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
22 Fourier Series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
23 Laplace Transform in Circuit Analysis . . . . . . . . . . . . . . . . . . . . . . 171
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
24 Fourier Transform in Circuit Analysis . . . . . . . . . . . . . . . . . . . . . . 181
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185
25 Second-Order Circuits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
26 Filters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
27 Wrapping Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211

Appendix. Solutions to Exercise Problems . . . . . . . . . . . . . . . . . . . . . . . 213


Bibliography (Some Textbooks Used in Colleges) . . . . . . . . . . . . . . . . . . 349
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
Voltage, Current, and Resistance
1

Electric circuits, such as the one shown in Fig. 1.1, consist of multiple connected
electrical components so that electrons can flow through a closed loop.

Fig. 1.1 An example of an


electric circuit, which is
represented using a circuit
diagram

In order to analyze and design electric circuits, we must first understand some
fundamental electrical quantities: voltage, current, and resistance.
Voltage is the difference in electric potential between two points in a circuit and
is measured in volts (V). A typical reference point is ground (GND), which is a point
we choose to be 0 V. However, it is also common to measure voltage across a
component, as can be seen in Fig. 1.2. When expressing voltage as a variable, we
usually use v.

+ Q -
Electrical
Component

Fig. 1.2 Voltage across an electrical component. The “+” and “ ” labels correspond to the
positive and negative terminals of the component

Voltage across an electrical component is measured with respect to the negative


terminal of the component instead of ground. This is equivalent to the voltage of the

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 1


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_1
2 1 Voltage, Current, and Resistance

positive terminal minus the voltage of the negative terminal, both with respect to
ground.
To gain an intuitive understanding of voltage, let us imagine that you are hiking
up and down a hill as shown in Fig. 1.3. We can consider the bottom of the hill to
have an altitude of 0 and measure the altitude with respect to the bottom of the hill.
As you go uphill, your altitude increases. As you go downhill, your altitude
decreases. In this scenario, the bottom of the hill is like ground while the voltage
is like altitude. Both the bottom of the hill and ground are reference points while both
altitude and voltage represent differences with respect to a reference point.

Fig. 1.3 A person hikes up


and down the hill, which
corresponds to changes in
altitude

Let us say that there is a meadow on the side of the hill like in Fig. 1.4. If we want
to determine the change in altitude of the meadow itself, we can measure the altitude
of the meadow with respect to the bottom of the meadow instead of the bottom of the
hill. This is akin to measuring the altitude at the top of the meadow, then subtracting
the altitude at the bottom of the meadow. If we consider the meadow to be like an
electrical component, the change in altitude of the meadow is like the voltage across
the component.

Fig. 1.4 The change in


altitude of a meadow
+

Change in
Altitude

Current is the flow of electrons through a circuit and is measured in amps (A).
We typically use i for current as a variable. In a closed loop, voltage causes current.
If there is no closed loop, current will not flow. Current through an electrical
component refers to the current flowing through that component.
1 Voltage, Current, and Resistance 3

We can view the relationship between voltage and current as a closed-loop water
system laid out on the hillside as shown in Fig. 1.5. There are two water tanks in
Fig. 1.5: one at the higher altitude, corresponding to higher voltage, and the other at
the lower altitude, corresponding to lower voltage. The water will naturally flow
from the upper tank to the lower tank through the water pipe. The altitude difference
h creates a gravitational force to push the water to flow from the upper tank to the
lower tank. Likewise, in an electric circuit, the voltage generates a pushing force to
drive the electric current. This water current corresponds to the electric current flow,
while the altitude difference h between the upper and lower tanks corresponds to the
voltage.

Fig. 1.5 A closed-loop water


system

Resistance is a measure of the material’s opposition to the flow of current and is


measured in ohms (Ω). Referring back to Fig. 1.5, the water pipes have friction that
inhibits the flow of water, which is similar to how resistance inhibits the flow of
electrons.
Voltage, current, and resistance are closely related to each other, and these
quantities change based on the type of electrical component. In order to consistently
analyze these components, electrical engineers use passive sign convention, a
method for assigning the positive and negative terminals of a component as well
as the direction of current.
In Fig. 1.6, an electrical component is labeled in accordance with passive sign
convention. The positive and negative terminals of the component can be arbitrarily
assigned, but the current direction must be from the positive terminal to the negative
terminal. It does not matter how you initially chose the positive and negative

Fig. 1.6 Voltage and current + -


for an electrical component
using passive sign convention Electrical
Component
4 1 Voltage, Current, and Resistance

terminals, but you must be consistent for the entire analysis. Even if your answer
contains a negative voltage or current, you may not have made a mistake; it just
means that the terminals or the current may have been opposite of what you initially
expected.
Until now, we have been using a generic representation of an electrical compo-
nent, so let us look into some basic circuit elements. Current–voltage characteristic
curves (I–V curves) represent the relationship between current and voltage for the
component and can help us better understand how the component operates.
A short circuit, also known as a wire, is used to connect other components. The
voltage across a wire is 0 V, while the current through a wire can be anything. The
resistance is 0 Ω (Fig. 1.7).

Fig. 1.7 Representation of a I


wire and its I–V curve
+ -

An open circuit is a disconnection in the circuit. The voltage across an open


circuit can be anything while the current through an open circuit is 0 A. The
resistance is infinite (Fig. 1.8).

Fig. 1.8 Representation of an I


open circuit and its I–V curve
+ -

A DC voltage source is a circuit element that provides a fixed voltage, such as


5 V or 9 V, across it. “DC” stands for “direct current”, which means that current only
flows in one direction and does not change. The voltage across a DC voltage source
is the voltage it is intended to provide while the current through it can be anything.
The internal resistance, the resistance inside of a component, of a DC voltage
source is 0 Ω (Fig. 1.9).
A DC current source is a circuit element that provides a fixed current through
it. The voltage across a DC current source can be anything while the current through
it is the current it is intended to provide. A DC current source’s internal resistance is
infinite (Fig. 1.10).
Resistors, as depicted in Fig. 1.11, are electrical components with set resistances,
such as 330 Ω, 1 k Ω, and 10 k Ω. For a resistor, current is proportional to voltage,
and we will further examine this relationship in Chap. 3. The resistance R in a
Exercise Problems 5

Fig. 1.9 Representation of a I


DC voltage source with
voltage v and its I–V curve +

Fig. 1.10 Representation of I


a DC current source with +
current i and its I–V curve

Fig. 1.11 Two representations of resistors. In this book, we will be using the one on the left, which
commonly used in the USA

resistor depends on the properties of the material, geometric shape of the resistor,
and sometimes temperature of the resistor. More properties of resistors will be
explored in Chap. 6.

Notes
If the conductor has resistance, electric voltage is required to force the electric
charges to move in one direction in a circuit, forming electric current. The
circuit must be a closed loop.
When using passive sign convention, make sure to stay consistent through-
out the whole problem.
The I–V curves here are for ideal circuit elements, which are
approximations of their real-world counterparts.

Exercise Problems

Problem 1.1 Either of the following two symbols represents a DC voltage source.
Here “V” is an abbreviation of “Volts.” “Volt” is a unit of voltage.
6 1 Voltage, Current, and Resistance

Fig. P1.1

Determine whether the following configurations of voltage sources are valid or


invalid. Why?

Fig. P1.2

Problem 1.2 The purpose of a voltage source in a circuit is to cause the current to
flow in a circuit. The flow of the electric current can be converted into something
useful to us. For example, the electric current running through a heating wire can
generate heat. The electric current running through a light bulb creates light. The
electric current running through an electric motor causes motion. Please comment on
the circuit shown whether this circuit is useful.
Exercise Problems 7

Fig. P1.3

Problem 1.3 Even though we do not see them in everyday life, there are such things
called “current sources.” The ideal current source provides constant current, regard-
less the rest of the circuit. The symbol for a current source is shown below. Here “A”
is an abbreviation of “Amperes.” “Amperes” is a unit of current.

Fig. P1.4

Determine whether the following configurations of current sources are valid or


invalid. Why?
8 1 Voltage, Current, and Resistance

Fig. P1.5

Problem 1.4 Determine whether the following circuits are valid.

Fig. P1.6

Problem 1.5 Draw a schematic for the flashlight circuit.

Solutions to Exercise problems are given in Book Appendix.


DC Power Supply and Multimeters
2

Let us suppose you are asked to build a circuit in Fig. 2.1, then to measure the
resistance of each resistor, the voltage across each resistor, and the current flowing
through the circuit.

Fig. 2.1 A circuit with two


50 Ω resistors and one 6 V
voltage source

First, we will need to get two 50 Ω resistors. For the voltage source, we will be
using a DC power supply, a device that can provide electrical power with
specifications on voltage and current. To connect the circuit, we will need a
breadboard and some wires. A breadboard, also known as a prototype board, is a
board with existing internal connections that is used for building circuits. The
breadboard we will be using in this example is a solderless breadboard, which
contains holes for plugging in the terminals of the components.
Figure 2.2 illustrates the breadboard’s internal connections, which can be thought
of as wires connecting the holes. The middle two columns of the breadboard are
connected horizontally, but not across the notch between these two columns. The
outer two columns, also known as the power rails, of the breadboard are connected
vertically and are typically used to connect to the power supply. By convention, the
red column connects to the positive terminal, while the blue column connects to the
negative terminal.
The final circuit for Fig. 2.1 is shown in Fig. 2.3. You will need to set up the DC
power supply by setting the voltage to the voltage you want to supply, which is 6 V

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 9


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_2
10 2 DC Power Supply and Multimeters

Fig. 2.2 A breadboard is shown on the left and its internal connections are shown on the right

To power supply RED (+)

DC Power Supply
50
V
A
- GND + - + - GND +
5V fixed
50

To power supply BLACK (-)

Fig. 2.3 The circuit from Fig. 2.1 is built on the board and connected to the DC power supply

in this example. For circuit protection, you should also set a limit for the current,
which will vary depending on the circuit.
In some cases, we need to use more than one DC power supply in a circuit, like in
Fig. 2.4. One possible way to build the circuit in Fig. 2.4 is shown in Fig. 2.5.
Your power supply panel layout may be different from the example here, so be
sure to read the instructions before you connect your circuit to the power. As an
example, we can build the circuit of Fig. 2.4 with a different kind of power supply as
shown in Fig. 2.6.
Now that we have built a circuit, let us measure the voltage across a resistor using
a multimeter, which is a device that can measure voltage, current, and resistance.
There are two types of multimeters: hand-held digital multimeters and desktop
digital multimeters, which can be seen in Figs. 2.7 and 2.8. No matter which type of
2 DC Power Supply and Multimeters 11

Fig. 2.4 A circuit with three


50 Ω resistors and two DC
power supplies

DC Power Supply
50
V
A
- GND + - + - GND + 50
5V fixed

50

Fig. 2.5 The circuit from Fig. 2.4 is built on the board and connected to two of the three outputs
from a DC power supply

multimeter you are using, you must plug two probes into two of the proper ports of
the multimeter unit in order to use it.
To measure the voltage across a resistor, you select the DC voltage measurement
mode by pushing the button labeled as “DC V,” connect the “Input V HI” (or “V” if
the label is just “V”) to one end of the resistor of interest, and connect the “LO”
(or “COM” if the label is “COM” in your multimeter) to the other end of the resistor.
This allows you to use the multimeter as a voltmeter, which measures the voltage
across two points in a circuit. When you make the measurement, you must leave the
power on. You can also use the voltmeter to measure the voltage across the power
source, with “Input V HI” to one terminal of the power supply and “LO” to the other
terminal.
Figures 2.9 and 2.10 show the setup for measuring voltage across the second
resistor in the circuit from Fig. 2.4.
To measure the current, depending on your multimeter, you may need to push a
button to select the DC current measurement mode, then follow the steps shown in
Fig. 2.11. This allows you to use the multimeter as an ammeter, which measures the
current through its two terminals. If you would like to measure the current through a
resistor, never connect the ammeter across the resistor or across a source! You must
first disconnect the circuit at a certain point. A correct connection is shown in
Figs. 2.12 and 2.13. If you make a mistake, you may send too much current through
the ammeter and blow the fuse.
12 2 DC Power Supply and Multimeters

Fig. 2.6 The circuit of Fig. 2.4 is powered by a different kind of DC power supply

Fig. 2.7 A hand-held digital


multimeter


V A

V A
OFF

V

A COM

Finally, to measure the resistance of a resistor, you must remove the resistor from
the circuit and measure it when the meter is at the resistance measurement mode as
shown in Fig. 2.14. You may need to adjust the range to get better precision.
2 DC Power Supply and Multimeters 13

Push this button before making the measurement


Connect to one
end of the resistor
Input
VΩ HI

LO
DC V DC I Ω
I
AC V AC I Connect to the
Digital Multimeter
other end of the resistor

Fig. 2.8 Use of a desktop digital multimeter to measure the voltage across a resistor

50
V HI

LO
50

Fig. 2.9 Breadboard setup to measure voltage across a resistor

Fig. 2.10 Circuit representation of Figure 2.9, where the component with a “V” is the voltmeter

Input
V HI

LO
DC V DC I 
I Insert into the
AC V AC I
Digital Multimeter circuit

Push this button

Fig. 2.11 Use of a desktop digital multimeter to measure current


14 2 DC Power Supply and Multimeters

Fig. 2.12 Breadboard setup


to measure current through the
circuit

Fig. 2.13 Circuit


representation of Fig. 2.12,
where the component with an
“A” is the ammeter

Input
VΩ HI
R
LO
DC V DC I Ω
I
AC V AC I
Digital Multimeter

Push this button

Fig. 2.14 To measure the resistance, connect a resistor across “Input Ω HI” and “LO”

Notes
To measure the voltage between two points, you can simply connect one probe
to one point and the other probe to the other point. Be sure that the multimeter
is at DC V voltage setting.
To measure the current at one point in the circuit, you must disconnect the
circuit at that point and then insert the two probes of the multimeter there to
re-connect the circuit.
To measure the resistance of a resistor, you need to remove the resistor
from the circuit. You can disconnect at least one end of the resistor from the
circuit. Never attempt to measure the resistance while the power of the circuit
is on, and both ends of the resistor are still connected in the circuit.
Exercise Problems 15

Exercise Problems

Problem 2.1 You are given a power supply and a circuit schematic shown. Suggest
three ways to connect the power supply to the 1 kΩ resistor.

Fig. P2.1

Problem 2.2 Identify the mistakes in using a multimeter.

(a) Trying to measure the voltage across the power supply.

Fig. P2.2
16 2 DC Power Supply and Multimeters

Fig. P2.3

(b) Trying to measure the current through the resistor.

Fig. P2.4

(c) Trying to measure the resistance of the resistor.

Solutions to Exercise problems are given in Book Appendix.


Ohm’s Law
3

Ohm’s law is the most popular and useful law for an electrical engineer and is a
must-know if you want to work with any electric circuit. Ohm’s law is a relationship
between the voltage v, current i, and resistance R for a resistor.
Let us revisit the water system analogy from Chap. 1 to set up an intuitive
understanding of Ohm’s law. In Fig. 3.1, if we move the upper water tank higher,
the altitude difference between the two tanks is increased and the water will flow
faster than before.

Higher tank
2h

h
Water flow
Faster water
Lower tank
flow

Fig. 3.1 When the altitude difference is increased, the water flow is also increased

As a side note, this water flow analogy is only a conceptual tool. In fact, this
analogy is not an accurate description of an electrical system because the water flow
speed varies with the altitude difference in a nonlinear relationship, while electrical
current varies with voltage in a linear relationship.
If the voltage v across the resistor is doubled, then the current i through the
resistor is doubled accordingly. The relationship between voltage and current is

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 17


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_3
18 3 Ohm’s Law

linear; in other words, the voltage v is directly proportional to the current i.


Therefore, there exists such a constant R such that
v ¼ Ri

and this constant R is the resistance. The linear relationship above is Ohm’s law.
We apply Ohm’s law to each resistor individually in a circuit, as shown in
Fig. 3.2.

Fig. 3.2 Representation of a


resistor with resistance R and
I
Ohm’s law, which describes +
its current–voltage
relationship Slope =

Example
If the voltage across the resistor is 10 V and the current through the resistor is
50 mA, what is the resistance of the resistor?

Solution
We can calculate the resistance of the resistor using Ohm’s law. We first rewrite
Ohm’s law to solve for R, then plug in the values provided.
v 10 V
R¼ ¼ ¼ 200 Ω:
i 50 mA

Notes
Ohm’s law implies that for a given resistance R for a resistor, if you supply
more voltage v across the resistor, you get more current i flowing through
it. This law v ¼ Ri is the foundation of electrical engineering.
Exercise Problems 19

Exercise Problems

Problem 3.1 Use Ohm’s law to calculate the current in the circuit.

Fig. P3.1
20 3 Ohm’s Law

Problem 3.2 According to the partial circuit shown, use Ohm’s law to calculate the
voltage across the resistor. You must use the voltage polarity and current direction
specified in the figure.

Fig. P3.2
Exercise Problems 21

Problem 3.3 You are given an electrical element without any labels. You connect
the element with a variable voltage source. You make some voltage/current
measurements as shown in the table. What most likely is this element?

V (volts) I (amperes)
10 5
0 0
10 5

Fig. P3.3

Problem 3.4 True or False?

(a) If you double the voltage across the resistor, the current through it doubles.
(b) If you double the voltage across the resistor, the current through it halves.
(c) If you halve the current through the resistor, the voltage across it doubles.
(d) If you halve the current through the resistor, the voltage across it halves.
(e) If you double the resistance of a resistor and keep the voltage across the resistor
unchanged, the current through the resistor doubles.
(f) If you double the resistance of a resistor and keep the current through the resistor
unchanged, the voltage across the resistor doubles.

Problem 3.5 The total human body in water is approximately 300 Ω. The electric
current over 10 mA is life threatening if the current runs through the heart
(10 mA ¼ 0.01 A). How much voltage in the water can be lethal?

Solutions to Exercise problems are given in Book Appendix.


Kirchhoff’s Voltage Law (KVL)
4

There are two Kirchhoff’s laws, both of which are based on one concept: conserva-
tion. In this chapter, we will use hiking as an analogy to build an intuition about
Kirchhoff’s voltage law (KVL) (Fig. 4.1).
Total al titude

Total al titude

Fig. 4.1 A cartoon depiction of KVL using skiing

It is a nice weekend and there are mountains close by, so you decide to take a
hike. You park your car at the trailhead parking lot and write down the altitude a0.
You can pick any trail to hike up. The requirement is that you must write down the
altitude every time you take a break. After you reach your destination, you can
choose any other trail to come back to your car. It is not a surprise that after you
return to the parking lot, your altitude reading is the same value as what you wrote
down at the beginning of your hiking trip. Otherwise, you are at the wrong parking
lot!
Let us assume that your altitude records are a0, a1, a2, and a3, reflecting the path
taken in Fig. 4.2. Let the altitude gain v at each hiking segment be

v 1 ¼ a1  a0 ,

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 23


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_4
24 4 Kirchhoff’s Voltage Law (KVL)

Fig. 4.2 The path of a


closed-loop hiking trip

a1
a2

a0

a3

v 2 ¼ a2  a1 ,

v 3 ¼ a3  a2 ,

v 4 ¼ a0  a3 :

It is important to follow the rule that the altitude gain is defined as the end point
altitude minus the starting point altitude for each segment.
Some of these v values are positive, and some are negative. A positive v value
implies that you hiked upwards at the hiking segment, while a negative v value
implies that you hiked downwards at that segment.
Now let us sum up these v values.

v1 þ v2 þ v3 þ v4 ¼ a1  a0 þ a2  a1 þ a3  a2 þ a0  a3 ¼ 0:

The sum of the altitude gains for all segments is zero for any closed-loop hiking
trip. The net altitude gain for the entire closed-loop hike is zero simply because the
end point and the starting point are the same point.
The above closed-loop hiking “law” holds if we replace the “altitude gain” at each
segment by the “altitude drop”, which is the starting point altitude minus the end
point altitude.
KVL can be applied to any electric circuit by replacing “altitude” with “electric
potential” or “voltage”. By KVL, in an arbitrary loop, or closed path, of any electric
circuit, the total sum of the voltage drops across each element is zero. KVL also
holds if you replace “voltage gain” with “voltage drop”, but you must be consistent
for the entire loop in concern. Do not mix them up.
Just like hiking where you must know the hiking direction, for a chosen electric
circuit loop, you need to select a direction, which can be clockwise or counterclock-
wise. You can imagine that the current in this loop flows in this chosen direction.
This imagined direction may be wrong, but it does not matter when you apply KVL.
4 Kirchhoff’s Voltage Law (KVL) 25

Example
Figure 4.3 shows a complicated DC circuit with many closed loops, with one
loop highlighted using thicker lines. Write the KVL expression for that loop
with the provided direction and labeling.

Fig. 4.3 Applying KVL to a randomly chosen loop

Solution
The loop direction has already been assigned as clockwise, and each element has
already been labeled with “+” and “” signs. We can assign each element a
“direction” according to the given “+” and “” signs, where the positive direction
is from “+” to “” following the loop direction. The negative direction is from “”
to “+” following the loop direction. For each element in the loop, we add its voltage
to the existing sum of the voltages if it is in the positive direction and subtract its
voltage if it is in the negative direction. Starting from v1, the KVL equation becomes:

v1 þ v2  vb  v3  va ¼ 0:

Instead, if we had set the positive direction to be from “” to “+” following the
loop direction, we would still have gotten a valid KVL equation. The key is to be
consistent.

Some people find it easier to divide the elements into two groups: elements in the
positive direction and elements in the negative direction. The resulting KVL equa-
tion can be set up by summing the voltages for each group, then setting them equal to
each other.
26 4 Kirchhoff’s Voltage Law (KVL)

Notes
Kirchhoff’s voltage law (KVL) is a law of “what goes up must come down”. In
any closed loop, the voltage can increase and decrease when traversing all of
the elements in the loop. The net voltage change must be zero in a loop.
This law is based on energy conservation.

Exercise Problems

Problem 4.1 Using the given voltage polarities, set up KVL equations for the
following circuits:

Fig. P4.1

Problem 4.2 In this problem, we will use a new current source called controlled
current source, whose symbol is a diamond with an arrow inside (see the figure
below). The symbol for a regular current source is a circle with an arrow inside. For
example, a controlled current source is as follows:

Fig. P4.2
Exercise Problems 27

Here “2i” indicate the value of this current source, and this value is two times the
current value i, which is defined elsewhere in the circuit.

Fig. P4.3

Find the current i in this circuit.

Problem 4.3 Set up the KVL equations for the following Wheatstone bridge circuit.

Fig. P4.4

Problem 4.4 Do not simplify the circuit. Use the KVL to solve for the current i in
the circuit.

Fig. P4.5
28 4 Kirchhoff’s Voltage Law (KVL)

Problem 4.5 Use KVL to verify if the following circuit is valid.

Fig. P4.6
Exercise Problems 29

Problem 4.6 Use KVL to verify if the following circuit is valid.

Fig. P4.7

Solutions to Exercise problems are given in Book Appendix.


Kirchhoff’s Current Law (KCL)
5

Kirchhoff’s current law (KCL) can be intuitively understood using the river
analogy (Fig. 5.1). Rivers sometimes merge and split into other branches, like in
Fig. 5.2.

Fig. 5.1 A cartoon depiction


of KCL using traffic

Since there is nowhere else for the water to go, the total amount of water flowing
into a region is equal to the total amount of water flowing out. For the example given
in Fig. 5.2, this means that

i1 þ i2 ¼ i3 þ i4 þ i5 ,

where the water currents are labeled as i1, i2, . . ., i5. Likewise, by KCL, the total
current entering a junction of an electric circuit is equal to the total current exiting

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 31


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_5
32 5 Kirchhoff’s Current Law (KCL)

i1 i3

i4

i5
i2

Fig. 5.2 Rivers merge and fork

that junction. A junction can either be a node or a supernode. A node is an


uninterrupted stretch of wire that connects two or more circuit elements, while a
supernode is a portion of the circuit that may contain multiple elements.

Example
Write the KCL expressions for the supernode and the node marked in Fig. 5.3
using i1, i2, i3, i4, and i5.

Fig. 5.3 KCL can be applied to a node or a supernode, which are marked with the dotted lines
Exercise Problems 33

Solution
For the supernode, we use KCL to set the total current entering the supernode equal
to the total current exiting the supernode, obtaining the expression:

i1 þ i3 ¼ i2 þ i4 :

For the node, we use the same process to get

i1 þ i5 ¼ i2 :

Notes
Kirchhoff’s current law (KCL) is based on the principle of conservation of
electric charge. The sum of the current entering a junction is equal to the sum
of the current leaving a junction.

Exercise Problems

Problem 5.1 Use KCL to verify if the following circuit is valid.

Fig. P5.1
34 5 Kirchhoff’s Current Law (KCL)

Problem 5.2 Use KCL to verify if the following circuit is valid.

Fig. P5.2

Problem 5.3 Set up the KCL equations for the following Wheatstone bridge circuit.

Fig. P5.3
Exercise Problems 35

Problem 5.4 Find the current i1 in the circuit shown in Fig. P5.4.

Fig. P5.4

Problem 5.5 This circuit model a transistor, which has many applications such as
amplifiers. Find ib.

Fig. P5.5

Solutions to Exercise problems are given in Book Appendix.


Resistors in Series and in Parallel
6

Being connected in parallel and in series are two common configurations to connect
components, which are shown in Fig. 6.1. Components that are connected in
parallel share the same voltage across each component, while components that are
connected in series share the same current through each component.

Fig. 6.1 The circuit on the left shows components connected in parallel with the same voltage v.
The circuit on the right shows components connected in series with the same current i. The rectangle
represents a generic circuit element

When we have resistors in parallel and in series, we can combine the resistors into
one resistor with an equivalent resistance, as illustrated in Fig. 6.2.
When we combine R1, R2, and R3 in parallel, we get the following expression for
Req:

1 1 1 1
¼ þ þ :
Req R1 R2 R3

When we combine R1, R2, and R3 in series, we get

Req ¼ R1 þ R2 þ R3 :

The general format of the above two expressions holds for different numbers of
resistors being connected, with only the number of terms being added together
differing.

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 37


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_6
38 6 Resistors in Series and in Parallel

Fig. 6.2 For both resistors in parallel (top) and resistors in series (bottom), we can combine
multiple resistors into one equivalent resistor Req

We can gain an intuitive understanding of the expressions for equivalent resis-


tance by looking at the expression for the resistance R.

L
R¼ρ ,
A
where ρ is the resistivity of the material, L is the length, and A is the cross-sectional
area, as labeled in Fig. 6.3.

Fig. 6.3 Resistance


R depends on the resistor’s
length L, resistivity ρ, and area
A

If we have three resistors with the same L, ρ, and A, then they will all have the
same resistance R. Connecting these three resistors in parallel, like in Fig. 6.4, would
be like forming a new resistor with area 3A, resulting in
6 Resistors in Series and in Parallel 39

Fig. 6.4 The effect of three


resistors in parallel

L R
Req ¼ ρ ¼ ,
3A 3
which is what we expected from the expression for resistors in parallel.
Connecting those three resistors in series, as shown in Fig. 6.5, would be like
forming a new resistor with length 3L, so

3L
Req ¼ ρ ¼ 3R,
A
which is what we expected from the expression for resistors in series.

Fig. 6.5 The effect of three


resistors in series

Notes
When the resistors are connected in series, the equivalent resistance is larger
than the largest one, and the equivalent resistance is the sum of all individual
resistance.

(continued)
40 6 Resistors in Series and in Parallel

When the resistors are connected in parallel, the equivalent resistance is


smaller than the smallest resistance, and its reciprocal is the sum of the
reciprocals of the individual resistances. The reciprocal of resistance is called
conductance.

Toll $10 Toll $10 Toll $10

Toll $10 Toll $15

Exercise Problems

Problem 6.1 Ten 1 kΩ resistors are connected in series, the total resistance is

(a) 10 Ω
(b) 100 Ω
(c) 1 kΩ
(d) 10 kΩ
(e) 100 kΩ
Exercise Problems 41

Problem 6.2 Ten 1 kΩ resistors are connected in parallel, the total resistance is

(a) 10 Ω
(b) 100 Ω
(c) 1 kΩ
(d) 10 kΩ
(e) 100 kΩ

Problem 6.3 Two resistors R1 and R2 are connected in series. The total resistance is
1 kΩ.

(a) R1 is less than 1 kΩ.


(b) R1 is larger than 1 kΩ.
(c) R1 is 500 Ω.
(d) R1 and R2 must have the same resistance.
(e) R1 and R2 must not have the same resistance.

Problem 6.4 Two resistors R1 and R2 are connected in parallel. The total resistance
is 1 kΩ.

(a) R1 is less than 1 kΩ.


(b) R1 is larger than 1 kΩ.
(c) R1 is 2 kΩ.
(d) R1 and R2 must have the same resistance.
(e) R1 and R2 must not have the same resistance.

Problem 6.5 Four resistors R1, R2, R3, and R4 are connected in parallel. They satisfy
the relationship: R1 ¼ R2 < R3 ¼ R4. The total resistance is 1 kΩ.

(a) R1 is less than 1 kΩ.


(b) R1 is less than 2 kΩ.
(c) R3 is less than 1 kΩ.
(d) R3 is less than 2 kΩ.
(e) R3 is less than 4 kΩ.
(f) R1 is larger than 4 kΩ.
(g) None of the above.

Problem 6.6 Ten resistors are connected in parallel, with Rn ¼ nΩ, for n ¼ 1, 2, . . .,
10.

(a) Rtotal ¼ 1 + 2 + 3 + 4 + 5 + 6 + 7 + 8 + 9 + 10 ¼ 55 Ω


(b) Rtotal ¼ 11 þ 12 þ 13 þ 14 þ 15 þ 16 þ 17 þ 18 þ 19 þ 10
1
¼ 2:9290 Ω
42 6 Resistors in Series and in Parallel

(c) Rtotal < 1 Ω


(d) Rtotal ¼ 1þ2þ3þ4þ5þ6þ7þ8þ9þ10
12345678910
Ω

Fig. P6.1

Problem 6.7 Find the total resistance for the resistor network shown in Fig. P6.1.
Each resistor in the network is 1 Ω.

Solutions to Exercise problems are given in Book Appendix.


Voltage Divider and Current Divider
7

Two circuits that are commonly used in electrical engineering are voltage dividers
and current dividers. Figure 7.1 shows a typical voltage divider consisting of
resistors in series, with the voltage across one resistor as the output voltage vout. A
resistor’s share of the total voltage vin is proportional to its resistance, so in the
configuration of Fig. 7.1,

R2
vout ¼ vin :
R1 þ R2

If instead we were measuring the voltage across R1 as the output voltage, we


would get

R1
vout ¼ vin :
R1 þ R2

Fig. 7.1 An example of a


typical voltage divider

In general, the expression for vout can be written as follows:

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 43


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_7
44 7 Voltage Divider and Current Divider

R
vout ¼ vin ,
Rtotal
where R is the resistance of the resistor that vout is measured across, and Rtotal is the
equivalent resistance of all of the resistors in series.
Figure 7.2 shows a typical current divider consisting of resistors in parallel, with
the current through one resistor as the output current iout. A resistor’s share of the
total current i is inversely proportional to its resistance, so for Fig. 7.2,

Req
iout ¼ i ,
R1
where Req is the equivalent resistance of all of the resistors in parallel.

Fig. 7.2 An example of a


typical current divider

The general expression for iout has the same form as the expression for Fig. 7.3,
but R1 is replaced with the resistor the output current is flowing through.
Another common circuit that is based on resistors is the Wheatstone bridge. A
Wheatstone bridge consists of four resistors and a galvanometer, as shown in
Fig. 7.3. A galvanometer, which corresponds to the circle with a tilted needle, is
a device that measures current. In Fig. 7.3, R1, R3, and R4 are known, while R2 is
variable, as indicated by the arrow through the resistor symbol.

Fig. 7.3 A configuration of a


Wheatstone bridge
7 Voltage Divider and Current Divider 45

We can redraw Fig. 7.3 in the configuration of Fig. 7.4 to clearly illustrate the
bridge between the R1–R2 branch and the R3–R4 branch containing the
galvanometer.

Fig. 7.4 Another way to


draw the Wheatstone bridge

When the current ig measured by the galvanometer is 0, the bridge is balanced.


We can determine the conditions under which the bridge is balanced by treating each
branch as a voltage divider. We can only use the voltage divider expression if the
current flowing through both resistors is the same, which is the case if there is no net
current flowing through the bridge. Otherwise, we cannot use the voltage divider
expression for both branches.
If there is a voltage difference between nodes a and b, current will flow, so we
must first set

va ¼ vb :

Using the voltage divider expression for both branches, we get

R2 R4
va ¼ v ,v ¼ v :
R1 þ R2 b R3 þ R4
After substituting, this simplifies to

R2 R4
¼ ,
R1 þ R2 R3 þ R4
R2 R3 þ R2 R4 ¼ R1 R4 þ R2 R4 ,

and finally,

R2 R3 ¼ R1 R4 :

This relationship between the resistances allows us to use the Wheatstone bridge
for a variety of applications, such as measuring an unknown resistance.
46 7 Voltage Divider and Current Divider

Notes
In a voltage divider, the resistor with the largest resistance gets the biggest
share of the total voltage. In a current divider, the resistor with the largest
resistance gets the smallest share of the total current.
When appropriate, using voltage dividers and current dividers to solve for
voltages and currents in a circuit is much more convenient than solving
equations.

Exercise Problems

Problem 7.1 In the circuit shown in Fig. P7.1, R1 ¼ 1 kΩ. Find the values of other
resistors.

Fig. P7.1
Exercise Problems 47

Problem 7.2 Find the voltage v in the circuit shown in Fig. P7.2. All resistors have
the value of 1 Ω.

Fig. P7.2

Problem 7.3 Calculate currents i1, i2, i3, and i4 in the circuit shown in Fig. P7.3.

(a) R1 ¼ R2 ¼ R3 ¼ R4 ¼ 1 Ω
(b) R1 ¼ 1 Ω, R2 ¼ 2 Ω, R3 ¼ 3 Ω, and R4 ¼ 4 Ω

Fig. P7.3
48 7 Voltage Divider and Current Divider

Problem 7.4 Find i1 and i2 in the circuit shown in Fig. P7.4.

Fig. P7.4

Problem 7.5 For a voltage divider circuit shown in Fig. P7.5. R1 is 1 Ω. The voltage
across R1 is 1 V. Is it possible to determine the source voltage v and the value of the
other resistor R2?

Fig. P7.5

Solutions to Exercise problems are given in Book Appendix.


Node-Voltage Method
8

The node-voltage method is a method that solves for the voltages at each essential
node in a circuit using KCL. An essential node is a node that connects to three or
more components. We can divide the method into four steps:

1. Label each essential node and the directions of each current exiting/entering a
node. Select one node as ground.
2. Write KCL equations for each essential node, excluding ground.
3. Write expressions for the currents in terms of the node voltages. For resistors,
these expressions will come from Ohm’s law.
4. Substitute the expressions for the currents into the KCL equations and solve.

Example
Find currents i1, i2, . . ., and i6 in the circuit shown in Fig. 8.1, where R1, . . .,
and R6 are known.

Solution

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 49


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_8
50 8 Node-Voltage Method

Fig. 8.1 Node-voltage method circuit

1. Figure 8.2 shows the completed first step. There are three essential nodes, which
we labeled as v1, v2, and ground. We choose the node with the most connections
to be ground, which is indicated by the symbol with three horizontal lines, to
simplify our calculations.
2. The KCL equations are as follows:

i2 þ i3 þ i4 ¼ 0,

i4 ¼ i5 þ i6 or  i4 þ i5 þ i6 ¼ 0:
3. Next, we need to write expressions for each of the currents used in the KCL
equations. v1, R2, the 10 V source, and R1 are all in series, so we can treat this
portion of the circuit like the 10 V source and R2 swapped places to write

v1 þ 10 V
i2 ¼ :
R1 þ R2
We directly use Ohm’s law to write the expressions for the other currents.

v1
i3 ¼ ,
R3
v1  20 V  v2
i4 ¼ ,
R4
v2  30 V
i5 ¼ ,
R5
8 Node-Voltage Method 51

Fig. 8.2 Labeled circuit from Fig. 8.1

v
i6 ¼ 2 :
R6
4. After substituting, we get the following:

v1 þ 10 V v1 v1  20 V  v2
þ þ ¼ 0,
R1 þ R2 R3 R4
v2 þ 20 V  v1 v2  30 V v2
þ þ ¼ 0,
R4 R5 R6
which simplifies to
     
1 1 1 1 20 V 10 V
þ þ v1 þ v2 ¼  ,
R1 þ R2 R3 R4 R4 R4 R1 þ R2
     
1 1 1 1 30 V 20 V
v þ þ þ v ¼  :
R4 1 R4 R5 R6 2 R5 R4

From here, you can solve the system of linear equations; however, you would
like. For this example, we will solve the system using matrix operations. We first
rewrite the equations in matrix form:
52 8 Node-Voltage Method

2 1 1 1 1 3 2 20 V 10 V 3
þ þ   
6 R1 þ R2 R3 R4 R4 v
7 1 6 R R1 þ R2 7
4 5 ¼4 4 5:
1 1 1 1 v 30 V 20 V
þ þ 2 
R4 R4 R5 R6 R5 R4
The solution is

2 1 1 1 1 31 2 20 V 10 V 3
  þ þ 
v1 6 R1 þ R2 R3 R4 R4 7 6 R4 R1 þ R2 7
¼4 5 4 5:
v2 1 1 1 1 30 V 20 V
þ þ 
R4 R4 R5 R6 R5 R4

5. For this particular problem, there is an additional step since we are asked to solve
for the currents instead of just the node voltages. To solve for the currents, we will
use the expressions we already wrote in Step 3 and plug in v1 and v2 from Step
4. We did not write an expression for i1, but i1 ¼ i2 due to KCL.

Example
Consider the circuit in Fig. 8.3. Find v1 and v2.

Fig. 8.3 Node-voltage method circuit with a supernode


8 Node-Voltage Method 53

Solution

1. The circuit in Fig. 8.3 already has one node labeled as ground and two other nodes
labeled as v1 and v2. The reason why v2 + 20 V is not labeled as v3 is because there
is only a voltage source between the two essential nodes, making it difficult to

Fig. 8.4 Labeled circuit from Fig.8.3

write KCL equations, so we consolidate the two nodes into one supernode. The
remainder of the labeling is done in Fig. 8.4.
2. The KCL equations are as follows:

i1 þ i2 þ i3 ¼ 0,

i3 þ i4 þ i5 þ i6 ¼ 0:
3. Using Ohm’s law, we can write

v1
i1 ¼ ,
20 Ω
v1
i2 ¼ ,
25 Ω
ðv1  25 VÞ  v2
i3 ¼ ,
30 Ω
v2
i4 ¼ ,
10 Ω
54 8 Node-Voltage Method

v2 þ 20 V
i5 ¼ ,
30 Ω
ðv þ 20 VÞ  ð15 VÞ
i6 ¼ 2 :
40 Ω
4. After substituting, the KCL equations become

v1 v ðv  25 VÞ  v2
þ 1 þ 1 ¼ 0,
20 Ω 25 Ω 30 Ω
v2  ðv1  25 VÞ v v þ 20 V ðv2 þ 20 VÞ  ð15 VÞ
þ 2 þ 2 þ ¼ 0:
30 Ω 10 Ω 30 Ω 40 Ω
After simplifying, we get

37v1  10v2 ¼ 250,

10v1 þ 57:5v2 ¼ 712:5,

and finally, v1  3.58 V and v2   11.77 V.


This concludes the second example, but if we want to find other electrical
quantities, we can solve for them using v1 and v2.

Notes
The node-voltage method provides a general guideline for analyzing circuits
by setting up KCL equations. When it is difficult to set up a KCL equation,
such as when there is only a voltage source between two nodes, use a
supernode instead.
Exercise Problems 55

Exercise Problems

Problem 8.1 Set up node equations for the circuit given in Fig. P8.1.

Fig. P8.1

Problem 8.2 Set up node equations for a circuit containing a controlled source.

Fig. P8.2
56 8 Node-Voltage Method

Problem 8.3 Set up the node equations for a circuit, in which a voltage source is
between the two nodes, and there are no resistors between these two nodes.

Fig. P8.3

Problem 8.4 Set up the node equations for a circuit, in which a controlled voltage
source is between the two nodes, and there are no resistors between these two nodes.

Fig. P8.4
Exercise Problems 57

Problem 8.5 Set up node equations for the circuit, where a voltage source is
between a node and the reference node.

Fig. P8.5

Solutions to Exercise problems are given in Book Appendix.


Mesh-Current Method
9

The mesh-current method solves for the currents through each mesh, a loop
without any inner loops, by setting up KVL equations. The mesh-current method
can be split into three steps:

1. Identify each mesh and select the direction of the current (counterclockwise or
clockwise) through each mesh. The direction you choose does not matter.
2. Set up a KVL equation for each mesh. Express the voltages in the KVL equations
in terms of the mesh currents; for resistors, this can be done using Ohm’s law.
3. Solve for the mesh currents.

Example
Find the currents iR1, iR2, . . ., and iR6 in the circuit shown in Fig. 9.1.

Fig. 9.1 Mesh-current method circuit

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 59


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_9
60 9 Mesh-Current Method

Solution
1. There are three meshes in the circuit, and we can set the direction of each mesh
current as counterclockwise. Figure 9.2 contains the labeled circuit.

Fig. 9.2 Labeled circuit from Fig. 9.1

2. We can set up a KVL equation for each mesh current as follows. Each term in the
equation represents the voltage drop across each element in the mesh.

R2 i1  10 V þ R1 i1 þ R3 ði1  i2 Þ ¼ 0,

20 V þ R4 i2 þ R3 ði2  i1 Þ þ R5 ði2  i3 Þ  30 V ¼ 0,

30 V þ R5 ði3  i2 Þ þ R6 i3 ¼ 0:

This simplifies to

ðR1 þ R2 þ R3 Þi1  R3 i2 þ 0i3 ¼ 10 V,

R3 i1 þ ðR3 þ R4 þ R5 Þi2  R5 i3 ¼ 50 V,

0i1  R5 i2 þ ðR5 þ R6 Þi3 ¼ 30 V:


3. To solve this system of equations, we can rewrite it in matrix form
2 32 3 2 3
R1 þ R2 þ R3 R3 0 i1 10
6 76 7 6 7
4 R3 R3 þ R4 þ R5 R5 54 i2 5 ¼ 4 50 5
0 R5 R5 þ R6 i3 30

to get the solution:


9 Mesh-Current Method 61

2 3 2 31 2 3
i1 R1 þ R2 þ R3 R3 0 10
6 7 6 7 6 7
4 i2 5 ¼ 4 R3 R3 þ R4 þ R5 R5 5 4 50 5:
i3 0 R5 R5 þ R6 30

4. We must take an additional step to solve for the currents through the resistors. The
currents through the resistors can be written as follows:

iR1 ¼ iR2 ¼ i1,

iR3 ¼ i2  i1 ,

iR4 ¼ i2 ,

iR5 ¼ i3  i2 ,

iR6 ¼ i3 :

Example
Find all three mesh currents in the circuit shown in Fig. 9.3.

Fig. 9.3 Mesh-current method circuit with a supermesh


62 9 Mesh-Current Method

Solution
1. The three meshes already have the directions of their mesh currents set as
clockwise. However, when there is a current source between two meshes i1 and
i2, it becomes difficult to write KVL equations. We can address this problem by
using a supermesh, the large loop enclosing both meshes, instead of the two
meshes i1 and i2. We indicate the supermesh with the dotted arrow in Fig. 9.3.
2. The KVL equations for the supermesh and the mesh corresponding to i3 are

10i1  20 V þ 30i3 ¼ 0,

30i3 þ 40ði3  i2 Þ þ 30ði3  i1 Þ  20 V ¼ 0,

respectively. To solve for three unknowns, we need another equation, which we


can get using the current source as a constraint.

i2  i1 ¼ 1 A:

After simplifying, we get

i1 þ 0i2 þ 3i3 ¼ 2,

3i1  4i2 þ 10i3 ¼ 2,

i1 þ i2 þ 0i3 ¼ 1:
3. Solving the system of equations gives us i1  0.06 A, i2  1.06 A, and
i3  0.65 A.
From here, we can solve for other electrical quantities using i1, i2, and i3 if
we want.

Notes
The mesh-current method provides a general guideline for analyzing circuits
by setting up KVL equations. When it is difficult to set up a KVL equation,
such as when there is a current source between two meshes, use a supermesh
instead.
Exercise Problems 63

Exercise Problems

Problem 9.1 Set up the mesh equations and solve for the mesh currents.

Fig. P9.1

Problem 9.2 This circuit contains a voltage-controlled voltage source. Set up the
mesh equations and solve for the mesh currents.

Fig. P9.2
64 9 Mesh-Current Method

Problem 9.3 This circuit contains a current source. Set up the mesh equations and
solve for the mesh currents.

Fig. P9.3

Since we do not know how to express the voltage drop across a current source, we
want to avoid current source in our mesh equations. Using a super mesh can avoid
the current sources. For this problem, a super mesh is indicated in Fig. P9.4 as a
dotted loop.

Fig. P9.4
Exercise Problems 65

Problem 9.4 What if the current source is a controlled current source?

Fig. P9.5

Problem 9.5 Application of the mesh equations, considering a special case of a


current course in the circuit.

Fig. P9.6

Solutions to Exercise problems are given in Book Appendix.


Circuit Simulation Software
10

Computer simulation software is crucial for circuit design because it can model a
circuit’s behavior, including its voltages and currents throughout the circuit, without
actually building the circuit itself. The software we will be looking at in this chapter
is Multisim, a popular software developed by NI for circuit analysis.
Let us go through an example to illustrate how to use Multisim from the very
beginning. If you are using a Windows computer, you can run Multisim by clicking
on Start ! All Programs ! National Instruments ! Circuit Design Suite xx.x. !
Multisim xx.x. This opens up a workplace, which is shown in Fig. 10.1. To create a
new schematic, click on File ! New ! Schematic Capture. To save the schematic,
click on File ! Save As. To open an existing file, click on File ! Open and select a
file.

Fig. 10.1 Multisim workplace

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 67


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_10
68 10 Circuit Simulation Software

To place components, click on Place ! Components. In the Select Component


Window, click on Group to select the components that you need. Click OK to place
the component on the schematic. If you click OK again, you can put another one of
the same component on the schematic. If you want to rotate a component, just right
click on it and select Rotate 90 .
Let us select two resistors and a DC voltage source. The resistors are in the
“Basic” group, while the DC sources are in the “Sources” group, which correspond
to Figs. 10.2 and 10.3, respectively. Now we have a 10-V DC voltage source and two
10 kΩ resistors, as shown in Fig. 10.4.

Fig. 10.2 Select a resistor

After selecting the components, we need to wire the components together. Click
on Place ! Wire to drag and place the wires so that the circuit looks similar to the
one in Fig. 10.5.
You can change the value of a component by double-clicking that component and
typing in the new value. Let us change the value of R1 to 5 kΩ. Finally, add the
ground, which is in the “Sources” group because all circuits must be grounded before
the circuit can be simulated. The final circuit should look something like the circuit
in Fig. 10.6.
10 Circuit Simulation Software 69

Fig. 10.3 Select a DC voltage source

0 1 2 3 4 5 6

R2
10k:
V1 R1
12V 10k:

Fig. 10.4 Some components are selected to build a circuit

Fig. 10.5 The circuit after


connecting the components
70 10 Circuit Simulation Software

Fig. 10.6 The final circuit

To simulate the circuit, click on Simulate ! Run, or just click on the green
triangle on the toolbar. Click on the red square on the toolbar to stop simulation.
Now we will take some measurements in the circuit by adding a voltmeter and an
ammeter. The multimeter icon is on the vertical toolbar to the right. Put two
multimeters in the workplace in the setup of Fig. 10.7 and double-click on each of
the multimeters to set one as the voltmeter and the other one as the ammeter.
Remember, you must disconnect the circuit and use the ammeter to reconnect the
circuit in order to measure the current.

Fig. 10.7 Two multimeters are used to measure the voltage and current

After running the simulation again, you will see the measured voltage and current
values in the multimeter displays shown in Fig. 10.8.
Exercise Problems 71

Fig. 10.8 The multimeters display the measurements

Notes
Circuit simulation software provides a means for visualizing a circuit’s behav-
ior without even building it, so it is highly recommended to simulate circuit
designs before building them order to catch major flaws. Use whichever
software suits the project’s needs the best.

Exercise Problems

Problem 10.1 Use Multisim to simulate a circuit shown to determine the node
voltage. You can choose any resistor values.

Fig. P10.1
72 10 Circuit Simulation Software

Problem 10.2 Chapter 15 discusses operational amplifiers (Op-Amps). Simulate an


inverting amplifier with a sinewave input. Change the power supply values to
observe any potential clipping in the output. The op-amp circuit is given in
Fig. P10.2.

Fig. P10.2

Problem 10.3 Operational amplifier circuits are normally designed to operate from
dual supplies, e.g., +9 V and 9 V. This is not always easy to achieve and therefore
it is often convenient to use a single ended or single supply version of the electronic
circuit design. Find a single supply op-amp circuit and use Multisim circuit.

Solutions to Exercise problems are given in Book Appendix.


Superposition
11

We see the effects of superposition in our daily life. For example, in Fig. 11.1, an
audio amplification system can have multiple inputs such as voice input, music
input, and so on. You can test your audio system by checking each voice input and
music input separately. When all inputs come in simultaneously, your system is able
to produce the mixed sound.

Fig. 11.1 A multiple-input


audio system can produce
mixed sound

The superposition principle also works for linear input/output systems. For
electric circuits, the inputs are the independent voltage and current sources, while
the outputs are the voltage or current measurements at other elements. An indepen-
dent source is a source whose output does not depend on other electrical quantities
in the circuit.
Using superposition to solve for the voltage or current at a given point in the
circuit can be broken up into three steps:

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 73


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_11
74 11 Superposition

1. Identify all of the independent sources. In circuit diagrams, an independent source


will be represented as a circle with markings on the inside to indicate whether it is
a voltage or current source.
2. Choose one independent source and zero out all of the other independent sources.
Zeroing out a voltage source means treating the voltage source as a short circuit
while zeroing out a current source means treating the current source as an open
circuit. Solve for the output at that point in the circuit. Repeat this step for all of
the independent sources.
3. Add together each of the outputs obtained from Step 2 to get the total output at the
given point.

During the process, do not remove any of the resistors or dependent sources. A
dependent source is a source whose output depends on other electrical quantities in
the circuit. A dependent source is also referred to as a controlled source. In circuit
diagrams, a dependent source looks like a diamond with markings on the inside to
indicate whether it is a voltage or current source. Voltage sources have “+” and “”
on the inside, while current sources have an arrow on the inside.

Example
In Fig. 11.2, use superposition to find the voltage v.

Fig. 11.2 Superposition


circuit

Solution
1. There are two independent sources, a 2-A current source and a 4-V voltage
source.
2. Let us start with the 2-A current source, so we will first zero out the voltage source
like in Fig. 11.3. The two resistors form a current divider, so the current through
each resistor is 1 A. Using Ohm’s law, we have

v1 ¼ ð1 AÞð10 ΩÞ ¼ 10 V:

Here, v1 is the value of v for the first case.


11 Superposition 75

Fig. 11.3 The circuit after


zeroing out the voltage source

Next, we will analyze the effect of the 4 V voltage source, so we will zero out
the current source, resulting in Fig. 11.4. The two resistors form a voltage divider,
and we get
 
1
v2 ¼ ð4 VÞ ¼ 2 V:
2
Here, v2 is the value of v for the second case.

Fig. 11.4 The circuit after


zeroing out the current source

3. Combining the results from the above step gives the final result of

v ¼ v1 þ v2 ¼ 8 V:

Notes
If a circuit has multiple independent sources, analysis based on superposition
looks at one independent source at a time with all other independent sources
zeroed out. The total effect is the summation of the results from each analysis
with one source acting alone.
Do not remove any dependent source at any time, and make sure to follow
the same positive and negative labeling throughout the entire analysis.
76 11 Superposition

Exercise Problems

Problem 11.1 A student uses the superposition principle to solve the voltage v. The
student’s answer is wrong. Please help this student to find the mistake. Let us start
with Fig. P11.1.

Fig. P11.1

There are two sources in the circuit.


Case 1: Let us first remove the source on the left, obtaining Fig. P11.2. It can be
verified that the solution is v ¼ 0.

Fig. P11.2

Case 2: Let us first remove the source on the right, obtaining Fig. P11.3. The left
two resistors are in parallel. Therefore, these two right resistors can be combined into
a 0.5-k resistor, as shown in Fig. P11.3.

This is a voltage divider, and we have v ¼ 2/3 V.


Combining the results from these two steps, the final answer is v ¼ 2/3 V.
However, the correct answer for v is 2 V.
Exercise Problems 77

Fig. P11.3

Fig. P11.4

Problem 11.2 Another student tries to use the superposition principle to solve for
the voltage v in a different problem. The student’s answer is wrong. Please help this
student to find the mistake. Let us start with Fig. P11.5.

Fig. P11.5

There are two independent sources in this circuit, and we will use the superposi-
tion principle to solve this problem.
78 11 Superposition

Case 1: We remove the left source. The circuit becomes Fig. P11.6.

Fig. P11.6

The 1 k resistor on the right has no effect in the circuit. The two resistors on the
left is a current divider. Each 1kΩ resistor on the left get 0.5A of current. Using
Ohm’s law, v ¼ 500 V.
Case 2: We remove the right source. The circuit becomes Fig. P11.7.

Fig. P11.7

After combining the right two 1 kΩ resistors into a 0.5-kΩ resistor, we obtain a
voltage divider. In this case, v ¼ 2/3 V.
According to the superposition principle, we combine these two answers and
obtain the final answer of

2 1 1
v¼  ¼ V:
3 2 6
However, the correct answer is
Exercise Problems 79

v ¼ 499:5 V:

What is wrong?

Problem 11.3 Use the superposition principle to solve for the voltage v. The circuit
is

Fig. P11.

Solutions to Exercise problems are given in Book Appendix.


Thévenin and Norton Equivalent Circuits
12

For a two-terminal electric system with no source, a simple resistor can be used as an
equivalent model. Very often we treat a complicated electric appliance, like a
microwave oven, a vacuum cleaner, or a toaster, as a resistor. Clearly, the microwave
oven is much more complicated than a simple resistor, but this resistor model is
sufficient for our task.
However, many systems have sources in them and cannot be modeled as resistors.
An audio system that can drive speakers is an example of such a system with sources.
Therefore, we need a different way to model the circuits that make up these systems
when creating equivalent circuits. An equivalent circuit is a simplified circuit that
still exhibits the same behavior of the original circuit.
When a complex circuit only consists of sources and resistors connected linearly,
we can represent it as a Thévenin equivalent circuit or a Norton equivalent circuit.
A Thévenin equivalent circuit consolidates a circuit into a voltage source and a
resistor connected in series, as illustrated in Fig. 12.1.

Fig. 12.1 The Thévenin equivalent circuit to model a general resistive circuit

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 81


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_12
82 12 Thévenin and Norton Equivalent Circuits

Finding the Thévenin equivalent circuit for a general resistive circuit requires two
steps:

1. Measure vTh across the terminals a and b.


2. Find RTh by either measuring the short-circuit current isc, as shown in Fig. 12.2, to
get RTh ¼ viTh
sc
or zeroing out all of the independent sources in the circuit followed
by Step 2(a) or Step 2(b).
(a) If there are no dependent sources, combine all of the resistors between
terminals a and b into one resistor with resistance RTh using equivalences
of resistors in series and parallel.
(b) If there are dependent sources, add an external testing voltage or current
source, as demonstrated in Fig. 12.3, then calculate RTh as

vT
RTh ¼ :
iT

Fig. 12.2 The Thévenin


resistance RTh is calculated
using the short-circuit current

Fig. 12.3 The resistance RTh


is evaluated by an external
testing source, with internal
independent sources removed
12 Thévenin and Norton Equivalent Circuits 83

Example
Find the Thévenin equivalent circuit for the circuit in Fig. 12.4, where the
dependent current source depends on the current through the 8 Ω resistor.

Fig. 12.4 A circuit with a dependent current source

Solution
1. We will find vTh by setting up a KCL equation for node a.

vTh v  24
þ 4 þ 3ix þ Th ¼0
8 2
Since

vTh
ix ¼ ,
8
 
vTh v v  24
þ 4 þ 3 Th þ Th ¼ 0,
8 8 2
and we get vTh ¼ 8 V.
2. To find RTh, we will use the test source method with KCL. After zeroing out the
independent sources and adding a test voltage source, we get the circuit in
Fig. 12.5. Never remove dependent sources!
The KCL equation for node a is

vT v
iT ¼ ix þ 3ix þ ¼ 4ix þ T :
2 2
In this case,
84 12 Thévenin and Norton Equivalent Circuits

vT
ix ¼ ,
8
 
vT v
iT ¼ 4 þ T ¼ vT:
8 2
Thus,

vT
RTh ¼ ¼ 1 Ω:
iT
We can use vTh and RTh for the final Thévenin equivalent circuit in Fig. 12.6.

Fig. 12.5 The independent sources are removed, and an external test source is attached

Fig. 12.6 The Thévenin


equivalent circuit for the
circuit in Fig. 12.4

Another type of equivalent circuit is the Norton equivalent circuit, which


consists of a current source and a resistor in parallel as shown in Fig. 12.7.
Finding the Norton equivalent circuit for a general resistive circuit requires two
steps:
12 Thévenin and Norton Equivalent Circuits 85

1. Short the terminals a and b to measure iNor.


2. Find RNor by either measuring the open-circuit voltage voc between terminals a
and b to get RNor ¼ ivNooc or zeroing out all of the independent sources in the circuit
followed by Step 2(a) or Step 2(b).
(a) If there are no dependent sources, combine all of the resistors between
terminals a and b into one resistor with resistance RNor using equivalences
of resistors in series and parallel.
(b) If there are dependent sources, add an external testing voltage or current
source, then calculate RNor as

vT
RNor ¼ :
iT
This process is very similar to finding the Thévenin equivalent circuit.

Fig. 12.7 A Norton equivalent circuit on the left and its corresponding Thévenin equivalent circuit
on the right

To understand the relationship between the two circuits in Fig. 13.1, let us find the
Thévenin equivalent circuit for the Norton equivalent circuit on the left. By Ohm’s
law, the voltage across the terminals a and b is iNorRNor. RTh for the Norton
equivalent circuit is the resistance after removing the independent current source,
so RTh is RNor. Thus, for Thévenin and Norton equivalent circuits that represent the
same original circuit,

RTh ¼ RNor and vTh ¼ iNor RNor :

To the outside world, the actual circuit and the equivalent circuit give the same
results when you do circuit analysis, but the equivalent circuit is much easier to
analyze. This is the purpose of using a Thévenin equivalent circuit or a Norton
equivalent circuit.
One can alternate between a Thévenin equivalent and a Norton equivalent to
simplify a circuit, as illustrated in the example below.
86 12 Thévenin and Norton Equivalent Circuits

Example
Find the Thévenin equivalent for the circuit in Fig. 12.8.

Fig. 12.8 A circuit with two sources

Solution
We will simplify the circuit from the left to the right using equivalent circuits and
equivalences for resistors in parallel and in series (Figs. 12.9, 12.10, 12.11, and
12.12).

Fig. 12.9 The Thévenin circuit is converted to the Norton circuit in the shaded area. Notice that the
voltage source of 10 V is converted into a current source of (10/R) A
12 Thévenin and Norton Equivalent Circuits 87

Fig. 12.10 Combining the two resistors into one and combining two current sources into one in the
shaded area
88 12 Thévenin and Norton Equivalent Circuits

Fig. 12.11 The Norton circuit in the shaded area is converted into the Thévenin circuit. Notice that
the current source of (10/R + 2) A is converted into a voltage source of (5 + R) V
12 Thévenin and Norton Equivalent Circuits 89

Fig. 12.12 Finally, the two resistors are combined into one resistor with a resistance 3R/2 for the
final Thévenin circuit

Notes
It is possible to simplify any linear circuit, no matter how complex, to an
equivalent circuit with just a single voltage source and a resistor in series or a
current source and a resistor in parallel connected to a load.
Before finding the equivalent circuit, you must first separate the load out so
that the circuit has an output port.

“They’re just equivalent. Neither one is the off brand.”

1 k: 2 k:

1 k:
90 12 Thévenin and Norton Equivalent Circuits

Exercise Problems

Problem 12.1 Find the Thévenin equivalent circuit of the circuit in Fig. P12.1.

Fig. P12.1

Problem 12.2 Find the Thévenin equivalent circuit of the circuit in Fig. P12.2.

Fig. P12.2
Exercise Problems 91

Problem 12.3 Use the testing source method to find the Thévenin resistance RTh in
Problem 12.2.

Fig. P12.3

Problem 12.4 Find the Norton equivalent circuit of the circuit in Fig. P12.4.

Fig. P12.4
92 12 Thévenin and Norton Equivalent Circuits

Problem 12.5 Find the Norton equivalent circuit of the circuit in Fig. P12.5, using
the step-by-step Thévenin/Norton conversion method.

Fig. P12.5

Solutions to Exercise problems are given in Book Appendix.


Maximum Power Transfer
13

In circuits, power is the rate at which energy changes and is measured in watts (W).
For an arbitrary component, the power P can be expressed as

P ¼ IV,

where I is the current through the component and V is the voltage across the
component, following passive sign convention. The component is dissipating
power if the power is positive and supplying power if the power is negative.
For resistors specifically, P ¼ IV can be rewritten using Ohm’s law into the
following forms:

V2
P¼ ,
R
P ¼ I 2 R:

Since the expression for power is always positive, we can see that resistors always
dissipate power. With loads that are often represented as resistors, more power
dissipated through the load corresponds to more perceived quantities such as louder
sound from a speaker. We may want to maximize these perceived quantities, so we
will need to find a way to maximize the power dissipated in the load or in other
words, to find the maximum power transfer. This process will require using the
Thévenin equivalent circuit.
Suppose that you have an audio amplifier and a speaker serving as the load, as
shown in Fig. 13.1. Here, we will pretend that the audio amplifier is a DC system and
everything is resistive, so the audio amplifier can be modeled by a Thévenin
equivalent circuit with RTh and vTh. We will also treat the speaker as a resistor.
What resistance should the speaker have in order to get the loudest sound?

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 93


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_13
94 13 Maximum Power Transfer

Fig. 13.1 Find the optimal resistance RLoad that obtains the maximum power possible

This equivalent circuit is a voltage divider. The received power by the load can be
calculated as
   
RLoad vTh
pLoad ¼ vLoad iLoad ¼ vTh  ,
RTh þ RLoad RTh þ RLoad

RLoad v2Th
pLoad ¼ :
ðRTh þ RLoad Þ2

To find the optimal value, we can take the derivative of pLoad with respect to
RLoad, and then set the derivative to zero. This procedure results in

RLoad ¼ RTh :

When RLoad is too small, the voltage across the load is too small, resulting in small
power. When RLoad is too large, the current through the load is too small, also
resulting in small power.
As a side note, for an AC (alternating current) circuit where the voltage and
current are not constant, all the components in the equivalent circuit are complex
numbers. The resistance R becomes impedance Z, which is the AC counterpart to
resistance. The notation for RLoad becomes ZLoad, and the notation for RTh becomes
ZTh. Maximum power transfer is achieved when the load impedance is the complex
conjugate of the source impedance:

Z Load ¼ Z Th :

Selecting the load impedance like this is called impedance matching.

Notes
The source internal resistance and the load resistance share the total power in
the form of a voltage divider. When the load resistance matches the source

(continued)
Exercise Problems 95

internal resistance, the load gets 50% of the power that the source provides.
Fifty percent of the total power is the maximum power that the load can get.
If the load resistance is larger than the matching value, the system is more
efficient, but the total current is smaller than the matching current and the load
gets less power.

Exercise Problems

Problem 13.1 Find the maximum power delivered to R in the circuit in Fig. P13.1
when R is set for maximum power transfer?

Fig. P13.1
96 13 Maximum Power Transfer

Problem 13.2 In Problem 13.1, let R ¼ 2.5 Ω. What is the power provided by the
1 V voltage source? What is the power provided by the 1 A current source? In the
circuit of Fig. P13.2, what percentage of the power delivered to the load R ¼ 2.5 Ω
by the two sources?

Fig. P13.2

Problem 13.3 As shown by the result of Problem 13.2, when the load resistance
equals to the Thevenin resistance, the percentage of the power delivered to the load
can be less than 50%. Use an example to explain this phenomenon.

Solutions to Exercise problems are given in Book Appendix.


Operational Amplifiers
14

An operational amplifier (op amp) is a component that amplifies the voltage


difference between two input terminals and is represented by the triangle symbol
in Fig. 14.1. As an amplifier, the output voltage v0 ¼ A(vp  vn), where A is the
amplifier gain, vp is the voltage at the noninverting input, and vn is the voltage at the
inverting input. The output voltage cannot be greater than the positive power supply
voltage or less than the negative power supply voltage. If the output voltage goes
beyond these two voltages, the output will get saturated and stay at the limiting
value. For an ideal amplifier, the gain A is virtually infinity, so even a slight
difference between the inputs outputs a rail voltage, which refers to the power
supply voltages, depending on which input is larger. This configuration is also
known as a comparator circuit.

Fig. 14.1 A complete op amp symbol on the left and a simpler op-amp symbol without the positive
and negative power supplies on the right

One golden rule that applies to all ideal op amps is that virtually no current enters
the input terminals. In other words, the input impedance is very high. Op amps also
have very low output impedance, which means that the output voltage does not
change much even with different loads. However, there is more to consider for op
amps in the commonly used configurations of negative feedback. An op amp in
negative feedback feeds the output voltage back to the inverting input terminal, with
an example being the left circuit in Fig. 14.2. Be careful not to connect the circuit like
the right circuit in Fig. 14.2 because the circuit will instead be in positive feedback,
resulting in an unstable circuit.

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 97


G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_14
98 14 Operational Amplifiers

Fig. 14.2 A special type of negative feedback configuration called an inverting amplifier on the
left and an unstable circuit on the right

This brings us to the golden rule that only applies to op amps in negative
feedback: for negative feedback, there is virtually no voltage difference between
the noninverting input and the inverting input. This is because the negative feedback
configuration reduces the voltage difference between the two input terminals to zero
while stabilizing the output. Using the golden rules appropriately makes circuit
analysis much simpler.
Let us analyze the left circuit in Fig. 14.2. Since no current enters the op amp at
the input terminals,

i f ¼ is :

Since there is no voltage difference between the input terminals and v+ is


connected to ground,

v ¼ vþ ¼ 0 V:

Ohm’s law yields

vs v
is ¼ and i f ¼ o :
Rs Rf

Thus,

vs v
¼ o,
Rs Rf

that is,

Rf
v0 ¼  v
Rs s:
The expression indicates that the input signal vs is inverted and amplified by a
R
factor of Rsf to get the output signal v0, which is why this circuit is called an inverting
amplifier.
14 Operational Amplifiers 99

Let us take a look at another example of negative feedback, the non-inverting


amplifier in Fig. 14.3.

Fig. 14.3 A basic non-inverting amplifier

Since the current flowing into the op-amp is zero, the two resistors form a voltage
divider. Since there is virtually no voltage drop between the input terminals, we have

v ¼ vþ ¼ vs :

The voltage divider relationship yields

Rs
v s ¼ vo :
R f þ Rs

Thus,
 
R f þ Rs Rf
vo ¼ vs ¼ vs 1 þ :
Rs Rs
 
R
The expression contains no minus sign and only amplifies vs by 1 þ Rsf , hence
why it is called a non-inverting amplifier.
Another circuit with an op amp in negative feedback is the voltage follower in
Fig. 14.4.
There is no voltage drop between v+ and v and vo is directly connected to v, so

vs ¼ v ¼ vo :
100 14 Operational Amplifiers

Fig. 14.4 A voltage follower

The output signal follows the input signal. Since op amps have a high input
impedance and low output impedance, voltage followers are typically used as
buffers to deliver more power to the load connected to the output.
Let us imagine an audio amplifier circuit as a Thévenin equivalent with a high
output impedance, say 80 kΩ, connected to a speaker as a low impedance load, say
8 Ω. Directly connecting the load results in the load only getting 0.01% of vTh
because the circuit acts like a voltage divider. However, if we insert a voltage
follower between the load and the audio amplifier as shown in Fig. 14.5, the load
can get 100% of vTh across it.

Fig. 14.5 A voltage follower is used to deliver to power to the load

Op amp circuits are not constrained to those discussed above, with an example
being Fig. 14.6. Using superposition for the circuit in Fig. 14.6, we can see the
following:
14 Operational Amplifiers 101

Fig. 14.6 This circuit calculates vo ¼ 13.5va  10vb  2.5vc

• If only vc is available, this is an inverting amplifier and the output is

100
 v:
40 c
• If only vb is available, this is also an inverting amplifier and the output is

100
 v
10 b:
• If only va is available, this is a non-inverting amplifier and the output is
   
100 100 108
1þ va ¼ 1 þ va ¼ v:
40j10 8 8 a

Therefore,

108 100 100


vo ¼ v  v  v ¼ 13:5va  10vb  2:5vc :
8 a 10 b 40 c

When the output of the op amp is connected with a load Rload, do not apply KCL at
the op amp output terminal because we do not know how much current is flowing out
from the op amp. For an ideal op amp, the output voltage vo is independent from the
load and a range of currents can flow out from the output. Remember not to short the
output to the ground!
102 14 Operational Amplifiers

Notes
Op amps are popular as building blocks in analog circuits since they can
perform many mathematical operations. They are also useful for their high
input impedance and low output impedance. The ideal output impedance is
zero, so 100% of the output power can be delivered to the load.
For all ideal op amps, since the input impedance is so high, the current
flowing into the non-inverting (+) pin and the inverting () pin are negligibly
small. It is only valid to assume that the difference in voltage between the
non-inverting (+) pin and the inverting () pin is negligibly small when the op
amp is in negative feedback.
If the op amp is in positive feedback, we cannot assume that that the
difference in voltage between the non-inverting (+) pin and the inverting ()
pin is negligibly small. A positive feedback op amp circuit is not totally
useless; it can be used to construct an oscillator.
Exercise Problems 103

Exercise Problems

Problem 14.1 Express the output voltage vout in terms of the inputs v1 and v2.

Fig. P14.1

Problem 14.2 Consider a current source as the inverting input. Find the current
running into the output terminal.

Fig. P14.2
104 14 Operational Amplifiers

Problem 14.3 The circuit shown in Fig. P14.3 can be thought of a current source.
Find the range of the load RL, in which the current in RL is a constant. What is the
value of this constant current?

Fig. P14.3

Solutions to Exercise problems are given in Book Appendix.


Inductors
15

Fig. 15.1 A photo of


inductors and the symbol for
inductors

An inductor is a coiled wire, but it does not behave like a straight wire at all. A photo
of three inductors and the inductor symbol are shown in Fig. 15.1. An inductor is
characterized by its inductance L, giving rise to a new relationship between its
voltage and current:

di
v¼L :
dt

Inductors have some special properties that resistors do not have, such as how
inductors can store energy in the form of a magnetic field. To obtain an expression
for the energy stored in an inductor, we begin with the following general expression
relating power ( p) to energy (w):

dw
p¼ :
dt
Power can also be related to voltage and current as

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 105
G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_15
106 15 Inductors

 
di
p ¼ vi ¼ L i:
dt

After integrating the above expressions with respect to time, we have

1
w ¼ Li2 :
2
Another property is that an inductor does not allow sudden changes in current. In
fact, there exists a quantity called the time constant which governs how slowly its
current changes. For inductors, the time constant τ is equal to RL . In response to a
switch action at t ¼ 0, the current through the inductor as a function of time can be
found using three important values:

1. The initial current value i(0).


2. The final current value i(1).
3. The time constant τ ¼ RL.

When calculating the time constant, we should use the R that is connected to the
inductor after the switch action. The general mathematical expression for i(t) is given
as
t
iðt Þ ¼ ið1Þ þ ½ið0Þ  ið1Þeτ , for t  0:

The following example illustrates how an inductor reacts to a switch action.

Example
Consider the circuit in Fig. 15.2. The inductor is ideal and there is no resistance
in the inductor. The switch has been closed for a long time. At t ¼ 0, the switch
suddenly opens. What happens to the current in the inductor?

Fig. 15.2 A circuit with an inductor


15 Inductors 107

Solution
While the switch is closed, the current source sends all 1 A current through the
inductor because there is no resistance through the inductor.

At t ¼ 0, the current source stops providing current to the inductor. The current in
the inductor will slowly change to a new value.
What is the new value? There is no source to the right of the switch and there is a
resistor connected to the inductor, so the resistor consumes electric energy stored in
the inductor and converts it into heat while no energy is being added to the inductor.
After a long transition time, the current through the inductor will approach zero, as
shown in Fig. 15.3.

Fig. 15.3 The curve of the current in the inductor vs. time

In this example, i(0) is 1 A and i(1) is 0 A. The R that is connected to the inductor
after the switch action is 10 Ω, so the time constant τ ¼ RL ¼ 10 Ω ¼ 0:2 s.
2H

Plugging these values into the general expression for i(t), we get

iðt Þ ¼ e5t A, for t  0:

We can also determine the inductance L from a typical inductor structure shown
in Fig. 15.4.

Fig. 15.4 A typical inductor


108 15 Inductors

N 2 μA
L¼ ,
l
where N ¼ the number of turns, l ¼ length, A ¼ cross-sectional area, and μ ¼ perme-
ability of the core.
The equivalent inductance for inductors in series, like in Fig. 15.5, can be
calculated as

Leq ¼ L1 þ L2 þ . . . þ LN:

Fig. 15.5 Inductors in series

Similarly, the equivalent inductance for inductors in parallel, like in Fig. 15.6, can
be calculated as

1 1 1 1
¼ þ þ ... þ :
Leq L1 L2 LN

Fig. 15.6 Inductors in parallel

Notes
The voltage–current relationship of an inductor includes a derivative rather
than following Ohm’s law. At very low frequencies, the inductors behave
almost like a short circuit while at very high frequencies, they behave almost
like an open circuit. If there is a resistor connected to the inductor, the current
through the inductor cannot change suddenly.

(continued)
Exercise Problems 109

Generally speaking, any conductor can will have some inductive properties
and can be viewed as an inductor, but typical inductors are made from a
cylindrical coil of conducting wire for increased inductance.

Exercise Problems

Problem 15.1 The switch closes at t ¼ 0. Find the inductor current iL as function
of time.

Fig. P15.1
110 15 Inductors

Problem 15.2 We use the same circuit as in Problem 15.1. We assume that the
switch has been closed for a long time. The switch opens at t ¼ 0. Find the inductor
current iL as function of time.

Fig. P15.2

Problem 15.3 The switch in the circuit in Fig. P15.3 has been closed for a long time
before opening at t ¼ 0. Find the inductor’s current iL and the inductor’s voltage vL
for t  0.

Fig. P15.3

Solutions to Exercise problems are given in Book Appendix.


Capacitors
16

Fig. 16.1 A photo of


capacitors and the symbol to
represent a capacitor

A capacitor, as shown in Fig. 16.1, consists of two conducting layers separated by


dielectric material, or in other words, an insulator. As a result, no DC current can
pass through a capacitor. When there is voltage across a capacitor, an electric field is
generated, causing positive charge to build up on one plate of the capacitor and
negative charge to build up on the other plate.
A capacitor can be characterized by a quantity called capacitance C with the
following relationship between its voltage and current:
Z
1
v¼ iðt Þdt:
C

Capacitors have some special properties that resistors do not have. For example,
the voltage across a capacitor cannot suddenly change. Capacitors also have a time
constant τ ¼ RC which governs how slowly its voltage changes.
We can also write an expression for the capacitor voltage curve in response to a
switch action, which is similar to the inductor current curve. Assuming that there is a
switch action at time t ¼ 0, the capacitor voltage curve is determined by three
important values:

1. The initial voltage value v(0).


2. The final voltage value v(1).
3. The time constant τ ¼ RC.

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 111
G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_16
112 16 Capacitors

We should use the R that is connected to the capacitor after the switch action to
calculate the time constant τ. The general mathematical expression v(t) for the
voltage across a capacitor is given as
t
vðt Þ ¼ vð1Þ þ ½vð0Þ  vð1Þeτ , for t  0:

Example
Consider the circuit in Fig. 16.2. The capacitor is ideal and there is no DC
current leaking through it. The switch has been closed for a long time. At t ¼ 0,
the switch suddenly opens. What happens to the voltages across each of these
two capacitors?

Fig. 16.2 A circuit with two capacitors

Solution
When the switch is closed for a long time, DC current does not flow through the
capacitors, so the capacitors at this time can be treated as open circuits. Therefore, we
can ignore the capacitors when finding the initial values. The resulting circuit
resembles a voltage divider with a 12 V source, giving us

20
v1 ð0Þ ¼ 12  ¼ 4 V,
20 þ 40
40
v2 ð0Þ ¼ 12  ¼ 8 V:
20 þ 40
16 Capacitors 113

After the switch is open, no new energy will be added to the capacitors, so the
energy stored in the capacitors will be consumed by the resistors, which will convert
the energy to heat. Eventually, the capacitor voltages will discharge to zero.

v1 ð1Þ ¼ v2 ð1Þ ¼ 0:

There are two time constants. The horizontal capacitor/resistor pair has a time
constant of

τ1 ¼ R1 C1 ¼ ð20 kΩÞð5 μFÞ ¼ 100 ms ¼ 0:1 s:

The vertical capacitor/resistor pair has a time constant of

τ2 ¼ R2 C2 ¼ ð40 kΩÞð1 μFÞ ¼ 40 ms ¼ 0:04 s:

Thus,
t
v1 ðt Þ ¼ 4e0:1 ¼ 4e10t V, for t  0,
t
v2 ðt Þ ¼ 8e0:04 ¼ 8e25t V, for t  0:

The capacitance C can be described by the following expression involving


physical characteristics of a capacitor:

A
C¼ε ,
d
where

C ¼ capacitance [F],
ε ¼ dielectric constant [N/A2],
A ¼ overlapping area [m2],
d ¼ gap [m] (Fig. 16.3).

Equivalent capacitance for capacitors in series can be calculated as (see Fig. 16.4)
follows:

1 1 1 1
¼ þ þ ... þ :
C eq C1 C 2 CN

Equivalent capacitance for capacitors in parallel is given as (see Fig. 16.5)


follows:

C eq ¼ C 1 þ C 2 þ . . . þ C N:
114 16 Capacitors

Fig. 16.3 A typical capacitor


consists of two conductors
and dielectric material in
between

Fig. 16.4 Capacitors in series

Fig. 16.5 Capacitors in parallel

Note that the expressions for series and parallel connections are opposite for
capacitors and inductors. The expression for series capacitors resembles the expres-
sion for parallel inductors. The expression for parallel capacitors resembles the
expression for series inductors.
The capacitor also can store energy (w). Power ( p) is related to energy by

dw
p¼ :
dt
Power can also related to voltage and current as
 
dv
p ¼ vi ¼ v C :
dt
16 Capacitors 115

After integration of the above expressions over time, we have w ¼ 12 Cv2 . This
voltage-generated energy is in the form of electric field.

• An inductor acts almost like a short circuit at DC (and very low frequency) and
open circuit at high-frequency AC.

• A capacitor acts almost like an open circuit at DC (and very low frequency) and
short circuit at high-frequency AC.

Notes
The capacitors do not follow Ohm’s law, while the voltage and current are
related by a derivative expression.
At very low frequencies, the capacitors behave almost like an open circuit.
At very high frequencies, they behave almost like a short circuit.
The voltage across the capacitors cannot change suddenly.
116 16 Capacitors

Exercise Problems

Problem 16.1 The switch closes at t ¼ 0. Find the inductor current iL as function
of time.

Fig. P16.1

Problem 16.2 We use the same circuit as in Problem 16.1. We assume that the
switch has been closed for a long time. The switch opens at t ¼ 0. Find the
capacitor’s voltage vC as function of time.

Fig. P16.2
Exercise Problems 117

Problem 16.3 The switch in the circuit in Fig. P16.3 has been closed for a long time
before opening at t ¼ 0. Find the capacitor’s voltage vC and the capacitor’s current iC
for t  0.

Fig. P16.3

Solutions to Exercise problems are given in Book Appendix.


Analysis of a Circuit by Solving Differential
Equations 17

When a circuit contains capacitors and/or inductors, we must use derivatives to relate
voltages and currents.
For a capacitor:

dvC ðt Þ
i C ðt Þ ¼ C :
dt
For an inductor:

diL ðt Þ
vL ð t Þ ¼ L :
dt

Fig. 17.1 An RC circuit

Example
Find the voltage vR after t ¼ 0.

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 119
G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_17
120 17 Analysis of a Circuit by Solving Differential Equations

Solution
After t ¼ 0, we set up a KVL equation for the RC circuit shown in Fig. 17.1 as

vR þ vC ¼ 12,

iC R þ vC ¼ 12,

dvC ðt Þ
RC þ vC ¼ 12,
dt
dvC ðt Þ 1
¼ ðv  12Þ,
dt RC C
dvC ðt Þ 1
¼ dt,
vC  12 RC
d½vC ðt Þ  12 1
¼ dt:
½vC  12 RC

We will integrate both sides. On the left-hand-side, we use the integration formula
Z
1
dx ¼ ln x þ Constant:
x

On the right-hand-side, we use the integration formula


Z
dx ¼ x þ Constant:

After integrating both sides, we have

1
ln ½vC ðt Þ  12 ¼  t þ Constant:
RC
or

vC ðt Þ  12 ¼ AeRCt ,
1
for t  0,

where A is a constant determined by the initial condition. The above expression can
be rewritten as

vC ðt Þ ¼ 12 þ AeRCt , for t  0:
1

Since vC(0) ¼ 0, A ¼ 12. Finally, we have


17 Analysis of a Circuit by Solving Differential Equations 121

Fig. 17.2 At t ¼ 0, the voltage across the resistor is discontinuous, while the voltage across the
capacitor is continuous

vC ðt Þ ¼ 12  12eRCt , for t  0:
1

Hence,

vR ðt Þ ¼ 12  vC ðt Þ ¼ 12eRCt , for t > 0,


1

and

vR ðt Þ ¼ 0, for t < 0:

Unlike a capacitor, the voltage across the resistor has a discontinuous jump at
t ¼ 0 from 0 V to 12 V. On the other hand, this voltage jump never happens for a
capacitor because the energy stored in the capacitor takes time to charge or discharge
(see Fig. 17.2).
Solving a differential equation is not an easy task in general. We will avoid
differential equations as much as possible. This is the main motivation that we will
use the phasor notation for sinusoidal steady-state analysis and use the Laplace
transform for non-constant and non-steady-state cases as we will discover later.
For a first-order differential equation concerning only the DC power and
switching actions, we can readily write down the solution for the capacitor voltage
or inductor current if we know the initial value, final value, and the time constant.

Notes
Even though inductors and capacitors do not follow Ohm’s law, KCL and
KVL are still valid and can still be used to set up equations. The equations now
are differential equations.
We can solve some simple differential equations. The solution of differen-
tial equations is a function of time (instead of a number).
122 17 Analysis of a Circuit by Solving Differential Equations

Given x2 + x + 1 = 0 (1)
Thus, x(x + 1) + 1 = 0
x + 1 = -1/x (2)
Eq. (1) also gives
“What is x + 1 = -x2
Combining the 2 equations above gives
wrong?” x2 = 1/x
x3 = 1
x=1
Substituting into (1) yields
3=0

Exercise Problems

Problem 17.1 Set up a node equation for the circuit in Fig. P17.1. Then express the
equation in terms of iL.

Fig. P17.1
Exercise Problems 123

Problem 17.2 Set up a differential equation for i1 + i2.

Fig. P17.2

Problem 17.3 Set up a differential equation for v1  v2.

Fig. P17.3

Solutions to Exercise problems are given in Book Appendix.


First-Order Circuits
18

Both RL and RC circuits are first-order circuits because their voltage and current
can be related by a first-order differential equation. In Fig. 15.2, there is one inductor
in the circuit; it is clearly a first-order circuit.
In Fig. 16.2, there are two capacitors in the circuit, which could be a second-order
circuit. However, when the circuit is in action after the switch is open, those
capacitors act independently. Therefore, the circuit in Fig. 16.2 can be decomposed
into two first-order circuits after the switch is open.
For all first-order circuits, we can use the general exponential charging/
discharging expression (see also Fig. 18.1)

Fig. 18.1 For a first-order system, the response is an exponential function depending on the initial
value, the final value, and the time constant. A smaller time constant gives a faster respond

t
xðt Þ ¼ xð1Þ þ ½xð0Þ  xð1Þeτ , for t  0,

to find the time response by using three important values:

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 125
G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_18
126 18 First-Order Circuits

the initial value x(0),


the final value x(1), and
the time constant τ.

If solving a differential equation is challenging for you, you can substitute the above
exponential function into the differential equation to see if it is really the solution.
You can convince yourself whether the solution is correct by verification. Later you
will learn the Laplace transform, which will convert a differential equation into an
algebraic equation. An algebraic equation is easier to solve than a differential
equation.

Notes
With simple switch actions, the first-order RC or RL circuits have an expo-
nential response for the capacitor voltage or the inductor current. This expo-
nential function is uniquely determined by the initial condition, the final
condition, and the time constant.
t
xðt Þ ¼ xð1Þ þ ½xð0Þ  xð1Þeτ , for t  0:

Exercise Problems

Problem 18.1 The input of an RC circuit is a periodic square pulse sequence. The
period is 2T. The time constant of the RC circuit is τ. The output signal is the
capacitor voltage vC. Find the output signal’s maximum value vmax and the minimum
value vmin.
Exercise Problems 127

Fig. P18.1

Problem 18.2 A student tries to solve a problem in his own way, and he does not
get the correct answer. Please help him to find the error. In the problem, the switch
has been closed for a long time. The switch opens at t ¼ 0. Find the capacitor’s
voltage iC as function of time.
128 18 First-Order Circuits

Fig. P18.2

The student’s solution:


After the switch has been closed for a long time, the capacitor acts like an open
circuit. Therefore,

iC ð0Þ ¼ 0:

At t ¼ 0, the switch opens. Now the circuit does not have any source, and the
capacitor will eventually discharge to 0. Thus,

iC ð1Þ ¼ 0:

Recall the general solution

iðt Þ ¼ ið1Þ þ ½ið0Þ  ið1Þet=τ , for t  0:

The student’s solution is

iC ðt Þ ¼ 0, for t  0:

Solutions to Exercise problems are given in Book Appendix.


Sinusoidal Steady-State (Phasor)
19

Fig. 19.1 A sine wave


with v ¼ Vm cos (ωt + φ) with
radian frequency ω ¼ 2πf or
frequency f ¼ T1

If the input of a linear circuit is a sine wave (see Fig. 19.1), the voltages and currents
everywhere in the circuit are sine waves with the same radian frequency ω, but they
may have different amplitudes Vm and phase angles φ.
Let us assume that the input source is a cosine function. We construct a complex
function, e jωt, by adding an imaginary part of a sine function. It happens that, in
mathematics, this complex function is so special in a linear differential equation. It is
an “eigenfunction.” This eigenfunction has a property that the derivative of it is a
constant ( jω) multiplication of it, and the integral of it is another constant  ωj
multiplication of it.
Therefore, derivative and integration can be performed by doing constant multi-
plication. No differential equations are needed. Since the eigenfunction e jωt appears
in every expression, and do not need to carry it along with us when we work on
equations. At the end, we convert the phasor notation back to the real world by
putting the ωt thing back in and discarding the imaginary part.
“Phasor”, Vm ∠ φ, is a shorthand notation for v ¼ Vm cos (ωt + φ) with
ωt omitted.

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 129
G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_19
130 19 Sinusoidal Steady-State (Phasor)

When the input is a sine wave, the circuit is an AC (alternating current) circuit.
You do not treat an inductor as a short circuit and a capacitor as open circuit
anymore. They are treated as “resistors” and characterized by a complex value
known as the impedance. A common symbol for the impedance is Z.
By using impedance and phasor notation, Ohm’s law works for capacitors and
inductors as well as the resistors.
For a resistor:

v ¼ R i:

Let

i ¼ I m cos ðωt Þ,

and then

v ¼ RI m cos ðωt Þ:

Phasor notation:

V ¼ Z I,
 
with Z ¼ R, I ¼ Im ∠ 0 , and V ¼ Vm ∠ 0 .
For an inductor:

di
v¼L :
dt
Let

i ¼ I m cos ðωt Þ:

Then

v ¼ ωLI m sin ðωt Þ ¼ ωLI m cos ðωt  90Þ:

Phasor notation:

V ¼ ðjωLÞI ¼ ZI,

with

Z ¼ jωL ¼ ωL∠90 :
jπ π π
,e 2 ¼ cos þ j sin ¼ 0 þ j ¼ j,
2 2
19 Sinusoidal Steady-State (Phasor) 131


∴j ¼ 1∠90

in the polar system.


For a capacitor:

dv
i¼C :
dt
Let

v ¼ V m cos ðωt Þ,

and then

i ¼ ωCV m sin ðωt Þ ¼ ωCV m cos ðωt  90Þ:

Phasor notation:

V
I ¼ ðjωC ÞV ¼
Z
with

1 1 
Z¼ ¼ ∠  90 :
jωC ωC
jπ π π 1
,e 2 ¼ cos þ j sin ¼0j¼ ,
2 2 j
1 
∴ ¼ 1∠  90
j
in the polar system.
What is a phasor anyway? A phasor Vm ∠ φ can represent v ¼ Vm cos (ωt + φ),
but they are not the same thing. A phasor Vm ∠ φ is a complex number with a real
part and an imaginary part, while v ¼ Vm cos (ωt + φ) is real.
Let us do some investigation to see how the phasor Vm ∠ φ is formed. Let us first
form a complex number that uses Vm cos (ωt + φ) as its real part and artificially adds
an imaginary part Vm sin (ωt + φ). This complex number can be expressed with an
exponential function:
jðωtþφÞ
V m cos ðωt þ φÞ þ jV m sin ðωt þ φÞ ¼ V m e ¼ V m ejωt ejφ :

We then throw away e jωt from the expression above, obtaining our phasor

V m ∠φ ¼ V m ejφ ,
132 19 Sinusoidal Steady-State (Phasor)

which is a complex number in the polar form (instead of the Cartesian form). By the
way,
pffiffiffiffiffiffiffiffi
j ¼ 1:

If you want to multiply two complex numbers, it is easier to use the polar form. If
you want to add two complex numbers, it is easier to use the Cartesian form. These
two forms can be converted to each other by using Euler’s formula:

ejx ¼ cos x þ j sin x:

When you use impedance Z and phasor notation for voltages and currents,
everything we discussed before works, including Ohm’s law, KVL, KCL, node-
voltage method, mesh-current method, Thévenin equivalent, and Norton equivalent.

Example
We use the circuit in Fig. 19.2 as an example to show how to Ohm’s law for
sinusoidal steady-state analysis. Find i(t) with

vs ðt Þ ¼ 80 cos ð2000t Þ V:

Fig. 19.2 An AC circuit with a capacitor, an inductor, and a resistor


19 Sinusoidal Steady-State (Phasor) 133

Solution
Write the voltage source in phasor notation as

V s ¼ 80∠0 :

The three “resistors” are in series and the total impedance is

1
Z ¼ R þ jωL þ ,
jωC
1
¼ 3000 þ jð2000Þð0:5Þ þ  ,
jð2000Þ 100  109

¼ 3000  j4000 Ω:

Ohm’s law gives

Vs
I¼ ,
Z

80∠0
¼ ,
3000  j4000

80∠0
¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ,
3000 þ 40002 ∠ tan 1 4000
2
3000

¼ 16∠53:13 mA:

Finally, change the phasor notation back to the normal time-domain notation
 
iðt Þ ¼ 16 cos 2000t þ 53:13 mA:

Example
Given that

is ðt Þ ¼ 10 cos ð50, 000t Þ A

and

vs ðt Þ ¼ 100 sin ð50, 000t Þ V,

find the steady-state expression of v(t) in the circuit of Fig. 19.3.


134 19 Sinusoidal Steady-State (Phasor)

Fig. 19.3 An AC circuit with two sources

Solution
First, we rewrite the sources in phasor notation and convert the capacitance and
inductance into impedance as shown in Fig. 19.4.

Fig. 19.4. The phasor notation is used for the circuit of Fig. 81

A node-voltage equation can be set up as



V V V V  100∠  90
10 þ þ  þ þ ¼ 0,
5 j 20
9
j5 20

V ¼ 10  j30 ¼ 31:62∠  71:57 V:

Therefore,
 
vðt Þ ¼ 31:62 cos 50, 000t  71:57 V:

Finally, we summarize this new concept of phasor.


When we do linear circuit analysis with sinusoidal inputs, we assume the steady
state, which means that the power is on for a long time and the system is stabilized.
In a linear system, the voltages and currents have the same frequency everywhere in
the circuit, except for their amplitudes and phases. The phasor short-hand notation is
a convenient way to represent the amplitude and phase, omitting the frequency
which never changes.
19 Sinusoidal Steady-State (Phasor) 135

The phasor transform is to add an imaginary part Vm sin (ωt + φ) to the real
function Vm cos (ωt + φ), resulting in a complex function

V m cos ðωt þ φÞ þ jV m sin ðωt þ φÞ,


jðωtþφÞ
¼ V me ,

¼ V m ejωt ejφ ,

and to drop to frequency part e jωt, obtaining a compact notation Vme jφ or Vm ∠ φ.


The inverse phasor transform is to write down the real function Vm cos (ωt + φ)
when you are given a complex number in the polar coordinate system Vm ∠ φ. You
can treat Vm ∠ φ as a normal complex number. You can multiply, divide, add, and
subtract Vm ∠ φ with other complex numbers. When doing the inverse phasor
transform, you must know the radian frequency ω because it is not given in the
phasor notation.
The concept of impedance is an extension of resistance, to be used with the
phasor notation. For an inductor, the impedance is

jωL ¼ ωL∠90 :

For a capacitor, the impedance is

1 1 
¼ ∠  90 :
jωC ωC

Notes
In sinusoidal steady-state analysis, the frequency never changes. The phasor is
a short-hand notation to represent the amplitude and phase of a sinusoidal
function. A phasor is also a complex number expressed in the polar coordinate
system.
Vm ∠ φ is a short-hand expression of Vm cos(ωt + φ),

V m ∠φ ¼ V m ejφ ¼ V m cos φ þ jV m sin φ:

The inductor’s impedance is jωL. The capacitor’s impedance is jωC


1
. When
using impedance, Ohm’s law works again.
136 19 Sinusoidal Steady-State (Phasor)

Exercise Problems with Solutions

Problem 19.1 Express the following signals in the phasor form:

(a) 5 cos (100t)


(b) 5 sin (100t)

(c) 5 cos (100t + 45 )

(d) 5 sin (100t + 45 )
(e) 2 cos (ωt)
(f) 2 cos (ωt)  3 cos (2ωt)
(g) 10
(h) 2t2 cos (ωt)

Problem 19.2 Express the transfer function in the phasor form. The input is vin and
the output is vC.

Fig. P19.1

Solutions to Exercise problems are given in Book Appendix.


Function Generators and Oscilloscopes
20

A function generator and an oscilloscope are essential for us to do steady-state


analysis of electric circuits. A function generator can produce periodic sine waves,
square waves, triangle waves, sawtooth waves, and so on. It can perform other
sophisticated tasks such as amplitude modulation. An oscilloscope can display the
signals and do some measurements about them. This chapter serves as a tutorial for
first-time users of these two pieces of equipment.
A function generator and an oscilloscope may look very similar. Figure 20.1
shows the front panel of a typical function generator. One way to tell a function
generator from an oscillator is that a function generator has a section “Function” as
shown in Fig. 20.1 that contains various function forms such as “Sine”, “Square”,
and so on. You need to have a BNC (Bayonet Neill–Concelman)-alligator-clips
cable with an impedance of 50 Ohms for the connection (see Fig. 20.2). Figure 20.3
shows how the alligator clips are connected to the circuit.

Fig. 20.1 A function generator

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 137
G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_20
138 20 Function Generators and Oscilloscopes

Fig. 20.2 A BNC to alligator-clip cable for the signal generator output

Fig. 20.3 Connection of the


function generator to your
circuit

For example, if you want to generate a sine wave of 1 kHz. Do the following:

1. Make sure that the power cord is plugged in. Push the on/off switch to power on
the equipment.
20 Function Generators and Oscilloscopes 139

2. If you decide to use Channel 1 output of your function generator, connect a BNC
cable to CH 1. The other end of the BNC cable can be connected to an oscillo-
scope to monitor the output of the function generator.

3. A sinewave can be selected by pushing the Sine button under “Function.” Select
the Continuous mode if it is not yet selected automatically when power is turned
on.
140 20 Function Generators and Oscilloscopes

4. Push the front-panel CH1 Output On button to enable the output from channel
1 if your signal generator has more than one outputs. Some function generators
only have one channel for output.

5. Use the oscilloscope auto-scaling function to display the sine waveform on the
screen. An oscilloscope looks like a function generator. One way to tell an
oscilloscope is that an oscilloscope has vertical and horizontal controls, as
shown in Fig. 20.5.

6. By pushing the Frequency/Period button you can get ready to set the frequency
of the signal that you would like to generate.
20 Function Generators and Oscilloscopes 141

7. The Frequency/Period button toggles between Frequency setting and Period


setting. You know the Frequency setting is selected if you see a small triangle on
the screen pointing to Freq.

8. To set or change the frequency value, use the keypad to type in the numerical
value and use the bezel buttons to specify the units.

For example, you enter a value “5” from the keypad for the frequency value. At
this time, the bezel menus will automatically change to Units. You can select the
unit of the frequency that you desired, say, MHz.
You can change the Phase, Amplitude, and Offset values in the similar way.
142 20 Function Generators and Oscilloscopes

9. Alternatively, you can change the values using the general-purpose turn-knob. By
pushing the arrow keys and then turning the knob, you can change a specific digit.

You can change to different waveforms by pushing different buttons on the


control panel.
The function generator can be modeled as a Thévenin equivalent circuit (see
Fig. 20.4), with RTh ¼ 50 Ω. The settings of the function generator assume that your
circuit matches the impedance RTh ¼ 50 Ω. If you set the output amplitude of the
signal generator as 5 V, the signal generator will produce a sine wave of 10 V. If your
circuit’s input impedance happens to be 50 Ω, you will get a 5 V sine wave. You are
getting twice the voltage displayed on the function generator at the output terminal
when the there is nothing connected to it. The voltage amplitude that you set on the
function generator may not be what your circuit actually gets, depending on your
circuit’s input impedance. You can use an oscilloscope to measure the actual
amplitude.

Fig. 20.4 The function generator has an output impedance of 50 Ω


20 Function Generators and Oscilloscopes 143

Fig. 20.5 An oscilloscope

The front panel of a typical oscilloscope is shown in Fig. 20.5. Special probes are
required to use the oscilloscope (see Fig. 20.6). The oscilloscope has two input
channels. We will use one channel to monitor the input signal, and the other channel
to monitor the voltage at the node connecting two resistors (see Fig. 20.7 for
connections).
You have many display options: Channel 1 only, Channel 2 only, both channels,
and so on. Coupling refers to the method used to connect an electrical signal from
your test circuit to the oscilloscope. DC coupling shows all components of an input
signal. AC coupling, on the other hand, blocks the DC component of the signal.
The horizontal axis is the time axis on the display. The “Time/Div” knob controls
the zoom scale on the time axis. The vertical axis is the voltage. The “Volts/Div”
knobs control the zoom scale on the voltage axis.
144 20 Function Generators and Oscilloscopes

Fig. 20.6 An oscilloscope


probe

Fig. 20.7 Both the signal generator and the oscilloscope are connected to the circuit

The scope probes are passive, and they do not contain any amplifiers. The 10
setting of the probe attenuates the signal 10:1.
An oscilloscope’s trigger function synchronizes the horizontal sweep at the
correct point of the signal, essential for clear signal characterization (see
Fig. 20.8). Trigger controls allow you to stabilize repetitive waveforms and capture
single-shot waveforms.
The trigger makes repetitive waveforms appear static on the oscilloscope display
by repeatedly displaying the same portion of the input signal. Imagine the jumble on
the screen that would result if each sweep started at a different place on the signal, as
illustrated in Fig. 20.9. The scope also has an auto mode, which can significantly
reduce your frustration trying to get a stable display. An oscilloscope has many other
useful functions. For example, you can move the cursor around to measure the time
delay between two different signals and then you can convert the time delay to phase
delay for your steady-state circuit analysis.
If this is your first time to use an oscilloscope, you can follow the procedure
below to display a signal on the scope.
20 Function Generators and Oscilloscopes 145

Fig. 20.8 Trigger level must be properly chosen within the signal’s voltage range

Fig. 20.9 An un-triggered display because the trigger level is selected too high

1. Connect the scope-probe to the input signal source, which can be a node in the
circuit you are diagnosing.
146 20 Function Generators and Oscilloscopes

2. Select the input channel (say, CH 1) that you connected in Step 1 by pushing the
corresponding button.

3. Press Auto-Set.
20 Function Generators and Oscilloscopes 147

4. Adjust the vertical position and scale using the front-panel knobs. The vertical
position knob moves your displayed signal up and down. The vertical Volts/Div
knob scales the amplitude of your displayed signal.

5. Adjust the horizontal position and scale using the front-panel knobs. The hori-
zontal Volts/Div knob stretches or shrinks horizontally the signal displayed. The
horizontal position knob determines the starting position of the signal displayed
relative to trigger time.
148 20 Function Generators and Oscilloscopes

Notes
A function generator can generate periodic signals with a specified frequency
and wave form. They are usually used as the sinusoidal power source for
steady-state studies.
An oscilloscope can display the wave forms picked up from a circuit; it can
be treated as a fancy graphic voltmeter. A proper triggering setup is required in
order to have a stable display of a periodic wave form.

Exercise Problems

Problem 20.1 In Fig. 20.1, an oscilloscope is directly connected to a signal


generator to verify the signal generated. We set the peak-to-peak voltage of a
sinewave to be 10 V. However, the oscilloscope shows a 20 V peak-to-peak. Is
there anything wrong?

Fig. P20.1

Problem 20.2 How to use an oscilloscope to estimate the time constant of a first-
order circuit?

Solutions to Exercise problems are given in Book Appendix.


Mutual Inductance and Transformers
21

Fig. 21.1 In a transformer, two separate coils are magnetically coupled

When two separate coils are magnetically linked together, emf (voltage) can be
generated by mutual inductance, M (see Fig. 21.1). The current in coil #1, i1, can

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 149
G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_21
150 21 Mutual Inductance and Transformers

induce voltage in coli #2. Similarly, the current in coil #2, i2, can induce voltage in
coli #1. They are related by

di1
v2 ¼ M ,
dt
and

di2
v1 ¼ M :
dt

The induced voltage is created by the derivative of the current in the other coil.
For DC current, the derivative is zero, and it will not induce any voltage. Therefore, a
transformer does not work for DC.
Dot convention for mutually coupled coils:

• When the current enters the dotted terminal in one coil, the polarity of induced
voltage in the other coil is positive at its dotted terminal.
• When the current leaves the dotted terminal in one coil, the polarity of induced
voltage in the other coil is negative at its dotted terminal.

Using the mesh-current method, we have two differential equations for the circuit
in Fig. 21.1:

di1 di
vg þ i1 R1 þ L1  M 2 ¼ 0,
dt dt
di2 di
i 2 R 2 þ L2  M 1 ¼ 0:
dt dt
We can use the phasor notation to transform the above two differential equations
into two algebraic equations:

V g þ I 1 R1 þ jωL1 I 1  jωMI 2 ¼ 0,

I 2 R2 þ jωL2 I 2  jωMI 1 ¼ 0:

In the matrix form, the system of equations becomes


    
R1 þ jωL1 jωM I1 Vg
¼ :
jωM R2 þ jωL2 I2 0

Thus,
21 Mutual Inductance and Transformers 151

   1  
I1 R1 þ jωL1 jωM Vg
¼ :
I2 jωM R2 þ jωL2 0

The solutions I1 and I2 are then transformed back to the time domain to get i1 and
i2. Finally, the voltage across the resistor R2 can be obtained via Ohm’s law.
In everyday life, an ideal transformer is a good approximation for a practical
transformer. A transformer is ideal if the coefficient of coupling k ¼ 1, where k ¼
pffiffiffiffiffiffiffi
M
and L1 ¼ L2 ! 1. In this ideal situation, the primary and secondary voltages
L1 L2
and currents are determined by the turn ratio, N1:N2, of the transformer.
In an ideal transformer (see Fig. 21.2), the magnitude of the volts per turn is the
same for each coil, or
   
 V 1  V 2 
  ¼  :
N 1  N 2 

The magnitude of the ampere-turns is the same for each coil, or

jI 1 N 1 j ¼ jI 2 N 2 j:

Fig. 21.2 Ideal transformer (phasor notation)

Even though the most popular application of transformers is to change AC


voltages, another important application is impedance matching, when the maxi-
mum power transfer is desired.
152 21 Mutual Inductance and Transformers

Fig. 21.3 A transform is used for impedance matching

Referring to Fig. 21.3 and considering an ideal transformer, we have

V 2 ¼ aV 1 ,

I 1 ¼ aI 2 :

The reflected impedance of ZL virtually on the primary side is

V1 1 V 1
Z reflected ¼ ¼ 2 2 ¼ 2 ZL:
I1 a I2 a
For example, if we have an audio amplifier with internal impedance Zs of 800 Ω
and a speaker ZL of 8 Ω, the speaker does not get much power if it is directly
connected to the audio amplifier output. The use of a 10:1 step-down transformer
(i.e., a ¼ 0.1) in between can greatly improve the power output. The maximum
power condition is

Z S ¼ Z reflected:

Notes
A basic transformer consists of two coils, sharing the same magnetic field. The
mutual inductance helps to induce an emf (voltage) in an adjacent coil. The
induced emf is proportional to the rate of change in the current. Therefore, a
transformer does not work for DC currents.
The turn ratio of a transformer determines the voltage ratio and current
ratio.
A transformer can also be used for impedance matching.
Exercise Problems 153

Exercise Problems

Problem 21.1 An ideal transformer has 1000 turns in its primary coil and 100 turns
in its secondary coil. Determine whether the following statements are true.

(a)This is a 10:1 transformer.


(b)This is a 1:10 transformer.
(c)This is a 1:0.1 transformer.
(d)This is a 0.1:1 transformer.
(e)This is a step-up transformer.
(f)
This is a step-down transformer.
(g)If the primary voltage is 10 V, the secondary voltage is 100 V.
(h)If the primary voltage is 10 V, the secondary voltage is 1 V.
(i)
If the primary current is 10 A, the secondary current is 100 A.
(j)
If the primary current is 10 A, the secondary current is 1 A.
(k)The transformer consumes power.
(l)
The transformer only works for a DC source input.
(m) The transformer only works for an AC source input.
(n)The frequency on secondary side is ten times higher than the frequency on the
primary side.
(o) The frequency on secondary side is ten times lower than the frequency on the
primary side.
(p) If the secondary side has a load of 100 Ω, the reflected impedance on the
primary side is 1000 Ω.
(q) If the secondary side has a load of 100 Ω, the reflected impedance on the
primary side is 10,000 Ω.

Problem 21.2 This problem about the dot notation and convention in a transformer.
Express the induced voltages for each case.

(a)

Fig. P21.1
154 21 Mutual Inductance and Transformers

(b)

Fig. P21.2

(c)

Fig. P21.3

(d)

Fig. P21.4
Exercise Problems 155

(e)

Fig. P21.5

(f)

Fig. P21.6

(g)

Fig. P21.7
156 21 Mutual Inductance and Transformers

(h)

Fig. P21.8

Solutions to Exercise problems are given in Book Appendix.


Fourier Series
22

We know how to do steady-state linear circuit analysis with sinusoidal inputs by


using phasor notation and impedance. What shall we do if the voltage source
produces a train of square waves or triangle waves, instead of sine waves? One
way to solve this problem is to convert the periodic non-sinusoidal waves into
sinusoidal waves. This procedure is called Fourier series expansion.

The theory of Fourier series expansion is that every periodic function can be
expressed as a combination of multiple sine waves

X
1
f ð t Þ ¼ a0 þ ½an cos ðnω0 t Þ þ bn sin ðnω0 t Þ,
n¼1

where f(t) ¼ f(t + T) is a periodic function with a period T,

ω0 ¼ 2π
T is the fundamental frequency,
2ω0, 3ω0, 4ω0, ⋯ are the harmonic frequencies,
and an and bn are the Fourier coefficients.

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 157
G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_22
158 22 Fourier Series

We now use an example to illustrate how to use the Fourier expansion method to
do steady-state circuit analysis with a non-sinusoidal input. Figure 22.1 gives a
periodic function f(t) with a period T. Assume that
f ðt Þ ¼ sin ðω0 t Þ þ sin ð3ω0 t Þ:

Its fundamental wave is a sine function sin(ω0t) with the same period T as f(t). Its
third harmonic wave sin(3ω0t) has a period of T/3. This function f(t) can be
represented by the superposition of two sine waves, one with the fundament fre-
quency and the other with the third harmonic wave.

Fig. 22.1 A periodic function f(t), its fundament wave, and its third harmonic wave

When doing circuit analysis, the two circuits in Fig. 22.2 are equivalent. You can
either use the source f(t) or use its expansion; you will get the same answer. Our goal
here is to find the voltage across the capacitor in the circuit.
There are two voltage sources in Fig. 22.2 (right), and these two sources have
different frequencies. We do not know how to handle two different frequencies
simultaneously when doing steady-state analysis with the phasor notation.
Our approach here is to use the superposition principle and to consider one
frequency at a time as shown in the circuits in Fig. 22.3 in the phasor notation.
The output voltage in both circuits in Fig. 22.3 can be found by using the voltage
divider.

 
1
jω0 C
V 1 ¼ 1∠  90 ,
R þ jω0 L þ 1
jω0 C

1
jω0 C
¼ j ,
R þ jω0 L þ 1
jω0 C

1
¼j ,
jω0 RC  ω20 LC þ 1
22 Fourier Series 159

 
1 
1 ω0 RC
¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2 ∠ 90  tan :
1  ω20 LC
1  ω20 LC þ ðω0 RC Þ2

Similarly,

 
1
j3ω0 C
V 2 ¼ 1∠  90 ,
R þ j3ω0 L þ 1
j3ω0 C
 
1 
1 3ω0 RC
¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 ffi ∠ 90  tan :
1  9ω20 LC
1  9ω20 LC þ ð3ω0 RC Þ2

After the inverse phasor transform, we have

vðt Þ ¼ v1 ðt Þ þ v2 ðt Þ,
1
¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
1  ω0 LC þ ðω0 RC Þ2
2

  
 ω0 RC
 cos ω0 t þ ∠ 90  tan 1
1  ω20 LC

1
þ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 ffi
1  9ω20 LC þ ð3ω0 RC Þ2
  
 3ω0 RC
 cos 3ω0 t þ ∠ 90  tan 1 ,
1  9ω20 LC
 
1 1 ω0 RC
¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 sin ω 0 t  tan
1  ω20 LC
1  ω20 LC þ ðω0 RC Þ2

Fig. 22.2 Since f(t) ¼ sin (ω0t) + sin (3ω0t), these two circuits are equivalent
160 22 Fourier Series

Fig. 22.3 The circuit of Fig. 22.2 (right) is converted into the phasor notation and is decomposed
into two circuits with two different frequencies

 
1 1 3ω0 RC
þ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 ffi sin 3ω 0 t  tan :
1  9ω20 LC
1  9ω20 LC þ ð3ω0 RC Þ2

Our next question is how to find the Fourier expansion when we are given a
general periodic function f(t). We need a set of formulas to find the Fourier
coefficients an and bn:
Z
an ¼ f ðt Þ cos ðnω0 t Þdt, n ¼ 0, 1, 2, . . . ,
T
Z
bn ¼ f ðt Þ sin ðnω0 t Þdt, n ¼ 1, 2, . . . ,
T
R R R t þT
where T means integration over one period T. That is, T ¼ t00 for any user
preferred value t0.
We now show you how these two formulas for an and bn are derived by using the
orthogonality method. Orthogonality means two vectors are perpendicular (i.e.,
90 ) to each other. One simple example is the x–y coordinate system, in which the x-
axis and y-axis are orthogonal to each other as seen in Fig. 22.4.
By inspection of Fig. 22.4, vector (0, 3) is orthogonal to vector (3, 0). Vector
!
(1, 1) is orthogonal to vector (2, 2). In mathematics, two vectors v 1 ¼ ðx1 , y1 Þ and
!
v 2 ¼ ðx2 , y2 Þ are orthogonal if their inner product (or, dot product) is zero, that is,

! !
< v 1 , v 2 >¼ x1 x2 þ y1 y2 ¼ 0:

We can readily verify that

< ð0, 3Þ, ð3, 0Þ >¼ 0  3 þ 3  0 ¼ 0

and

< ð1, 1Þ, ð2, 2Þ >¼ 1  ð2Þ þ 1  2 ¼ 0:


22 Fourier Series 161

Fig. 22.4 The x–y coordinate


system

The inner product can be extended for two real one-dimensional (1D) functions
f and g by using the integration-defined inner product as
Z
< f , g >¼ f ðt Þgðt Þdt:

Two functions f and g are orthogonal if their inner product is 0. If the two
functions f and g are periodic with a period T, the integration interval is from t0 to
t0 + T, for an arbitrary value t0.
It is not difficult to verify that cos(nω0t) and cos(mω0t) are orthogonal, and sin
(nω0t) and sin(mω0t) are orthogonal, if n 6¼ m. We also have cos(nω0t) and sin(mω0t)
are orthogonal, for all n and m. In fact,
Z t 0 þT
dt ¼ T:
t0
Z t 0 þT
sin ðnω0 t Þdt ¼ 0, for all integers n:
t0
Z t 0 þT
cos ðnω0 t Þdt ¼ 0, for all non  zero integers n:
t0
Z t 0 þT
cos ðnω0 t Þ sin ðmω0 t Þdt ¼ 0, for all m and n:
t0

Z (
t 0 þT 0, for m 6¼ n
cos ðnω0 t Þ cos ðmω0 t Þdt ¼ T
t0 , for m¼n
2
162 22 Fourier Series

Z (
t 0 þT 0, for m 6¼ n
sin ðnω0 t Þ sin ðmω0 t Þdt ¼ T
t0 , for m ¼ n:
2
To find a0, we integrate both sides of

X
1
f ð t Þ ¼ a0 þ ½an cos ðnω0 t Þ þ bn sin ðnω0 t Þ
n¼1

over T:
Z t 0 þT Z t 0 þT
f ðt Þdt ¼ a0 dt,
t0 t0

1 Z
X t 0 þT
þ ½an cos ðnω0 t Þ þ bn sin ðnω0 t Þdt,
n¼1 t0

Z t 0 þT
f ðt Þdt ¼ Ta0 þ 0:
t0

Thus,
Z t 0 þT
1
a0 ¼ f ðt Þdt:
T t0

To find an, we multiply both sides of

X
1
f ðt Þ ¼ a0 þ ½an cos ðmω0 t Þ þ bn sin ðmω0 t Þ
m¼1

with cos(nω0t) and then integrate both sides over T:


Z t 0 þT
f ðt Þ cos ðnω0 t Þdt,
t0
Z t 0 þT
¼ a0 cos ðnω0 t Þdt
t0

1 Z
X t 0 þT
þ ½an cos ðmω0 t Þ cos ðnω0 t Þ þ bn sin ðmω0 t Þ cos ðnω0 t Þdt,
m¼1 t0
22 Fourier Series 163

Z t 0 þT
T
f ðt Þ cos ðnω0 t Þdt ¼ 0 þ a þ 0:
t0 2 n

Thus,
Z t 0 þT
2
an ¼ f ðt Þ cos ðnω0 t Þdt:
T t0

To find bn, we multiply both sides of

X
1
f ðt Þ ¼ a0 þ ½an cos ðmω0 t Þ þ bn sin ðmω0 t Þ
m¼1

with sin(nω0t) and then integrate both sides over T:


Z t 0 þT
f ðt Þ sin ðnω0 t Þdt
t0
Z t 0 þT
¼ a0 sin ðnω0 t Þdt
t0

1 Z
X t 0 þT
þ ½an cos ðmω0 t Þ sin ðnω0 t Þ þ bn sin ðmω0 t Þ sin ðnω0 t Þdt,
m¼1 t0

Z t 0 þT
T
f ðt Þ sin ðnω0 t Þdt ¼ 0 þ 0 þ b
t0 2 n:

Thus,
Z t 0 þT
2
bn ¼ f ðt Þ sin ðnω0 t Þdt:
T t0

Before we proceed to find a Fourier expansion of a given periodic function, we


have some short cuts.

1. If f(t) has no DC offset, a0 ¼ 0.


2. If f(t) is an even function, then all bn ¼ 0 because f(t) can only be built with even
functions cos(nω0t).
3. If f(t) is an odd function, then all an ¼ 0 because f(t) can only be built with odd
functions sin(nω0t).  
4. If f(t) has half-wave symmetry, that is, f ðt Þ ¼ f t  T2 [Delay T/2, becomes
negative], then even coefficients are zero. That is,

an ¼ bn ¼ 0, for n even:

An example of half-wave symmetry is shown in Fig. 22.5.


164 22 Fourier Series

Fig. 22.5 A function with


half-wave symmetry

5. As a special case of (3), if the half-wave itself is symmetric as indicated in


Fig. 22.6, the Fourier coefficient integral can be evaluated only over T/4:
Z T
8 4
an ¼ f ðt Þ cos ðnω0 t Þdt,
T 0
Z T
8 4
bn ¼ f ðt Þ sin ðnω0 t Þdt:
T 0

Fig. 22.6 A function with


half-wave symmetry

Example
Find the Fourier expansion for the periodic function defined in Fig. 22.7.
22 Fourier Series 165

Fig. 22.7 A periodic triangle


function f(t)

Solution
This function satisfies conditions (1), (3), (4), and (5). All we need to find is bn for
n being odd. We can write down the equation for the straight-line segment of f(t) on
[0, T/4] as
h i
4 T
f ðt Þ ¼ t, t 2 0, :
T 4
Then
Z T
8 4
bn ¼ f ðt Þ sin ðnω0 t Þdt,
T 0
Z T
8 4
4
¼ t sin ðnω0 t Þdt,
T 0 T
 T
32 sin ðnω0 t Þ t cos ðnω0 t Þ  4 2π
¼ 2   , with ω0 ¼ T ,
T n2 ω20 nω0 0

8 nπ
¼ sin , with n being odd:
π 2 n2 2
The Fourier series for f(t) is given as

8 X
1
1 nπ
f ðt Þ ¼ sin sin ðnω0 t Þ,
π2 n¼1, 3, 5, ...
n2 2
h i
8 1 1 1
¼ sin ð ω 0 t Þ  sin ð 3ω 0 t Þ þ sin ð 5ω 0 t Þ  sin ð7ω 0 t Þ þ . . . :
π2 9 25 49
166 22 Fourier Series

Last but not least, the Fourier expansion can also be used to find the power of the
function by using the Parseval’s theorem:
Z
1 X 2
T 1 
1
½ f ðt Þ2 dt ¼ a20 þ a þ b2n :
T 0 2 n¼1 n

The proof of the Parseval’s theorem is rather straight forward. The starting point
is the expansion

X
1
f ð t Þ ¼ a0 þ ½an cos ðnω0 t Þ þ bn sin ðnω0 t Þ:
n¼1

We square both sides and then integrate both sides over T. Using the orthogonal-
ity, the integration results are

Z Z ( X
1
2
½ f ðt Þ dt ¼ a0 þ ½an cos ðnω0 t Þ þ bn sin ðnω0 t Þg2 dt,
T T n¼1
Z Z
½ f ðt Þ2 dt ¼ a20 dt
T T

1  Z
X Z
þ an cos ðnω0 t Þdt þ bn sin 2 ðnω0 t Þdt ,
2 2 2

n¼1 T T

Z
T X 2
1 
½ f ðt Þ2 dt ¼ a20 T þ an þ b2n :
T 2 n¼1

Thus,
Z
1 X 2
T 1 
1
½ f ðt Þ2 dt ¼ a20 þ a þ b2n :
T 0 2 n¼1 n

Notes
Any periodic function can be represented as a combination of sine waves with
different frequencies. The period of the function determines the fundamental
frequency, which is the lowest frequency in the Fourier expansion. Other
frequencies (called harmonics) are positive integer multiples of the fundamen-
tal frequency.
The harmonic functions are orthogonal to each other. Fourier expansion
coefficient formulas are derived based on the orthogonality properties.
Exercise Problems 167

Exercise Problems

Problem 22.1 Match a Fourier series with a periodic function with ω0 ¼ 2π/T.

X
1
f 1 ð t Þ ¼ a0 þ an cos ðnω0 t Þ,
n¼1

X
1
f 2 ðt Þ ¼ an cos ðnω0 t Þ
n¼1

X
1
f 3 ð t Þ ¼ a0 þ bn sin ðnω0 t Þ,
n¼1

X
1
f 4 ðt Þ ¼ bn sin ðnω0 t Þ,
n¼1

X
1
f 5 ðt Þ ¼ ½an cos ðnω0 t Þ þ bn sin ðnω0 t Þ,
n¼1
n ¼ odd
168 22 Fourier Series

X
1
f 6 ðt Þ ¼ an cos ðnω0 t Þ,
n¼1
n ¼ odd

X
1
f 7 ðt Þ ¼ bn sin ðnω0 t Þ,
n¼1
n ¼ odd

X
1
f 8 ð t Þ ¼ a0 þ ½an cos ðnω0 t Þ þ bn sin ðnω0 t Þ,
n¼1

X
1
f 9 ðt Þ ¼ ½an cos ðnω0 t Þ þ bn sin ðnω0 t Þ:
n¼1

(a)

Fig. P22.1

(b)

Fig. P22.2

(c)

Fig. P22.3
Exercise Problems 169

(d)

Fig. P22.4

(e)

Fig. P22.5

(f)

Fig. P22.6

(g)

Fig. P22.7
170 22 Fourier Series

(h)

Fig. P22.8

(i)

Fig. P22.9

Problem 22.2 Show that the Fourier series has an equivalent exponential form

X
1
f ðt Þ ¼ cn ejnω0 t :
n¼1

Solutions to Exercise problems are given in Book Appendix.


Laplace Transform in Circuit Analysis
23

Why do we need this Laplace transform? The reason is that we do not want to solve
differential equations when the circuit contains capacitors and/or inductors.
Capacitors and inductors are troublemakers, making mathematical work harder
than necessary. We already have some methods to deal with them. The Laplace
transform method is just an alternative one.
For the first-order system, if the power sources are DC with some switch actions,
the first-order differential equation is easy to solve, and the solution is essentially an
exponential function that can be uniquely determined by three important values: the
initial value, the final value, and the time constant.
If the sources are sine waves and the steady-state solutions are to be sought, the
situation is fairly easy if the phasor notation is used. This phasor approach only
works for sinusoidal sources and for steady-state analysis. What if the sources are
periodical, but not exactly sinusoidal? This is where the Fourier series expansion
comes in. Every periodic function can be represented by a Fourier series
X1
f ð t Þ ¼ a0 þ n¼1
½an cos ðnω0 t Þ þ bn sin ðnω0 t Þ:

In fact, this equation does not hold at the points where f(t) is not continuous. At
the jumping points, the Fourier series converges to the average of the left and right
limits of f(t) at those points. In this tutorial, let us do not worry about those
discontinuities. Once the Fourier series expansion is obtained, the phasor method
can apply by using the superposition principle, considering one sine wave at a time.
What is the Laplace transform good for? It is good for the sources that are turned
on at t ¼ 0. Obviously, it will not work for any steady-state analysis. The Laplace
transform converts a real function f(t), t  0 to a complex function F(s), where s is a
complex number. The derivative of f(t) will correspond to sF(s)  f(0), where f(0)
is the initial condition of the function f(t) right before the source is turned on at t ¼ 0.
We have some important Laplace transform properties listed in Table 23.1.

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 171
G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_23
172 23 Laplace Transform in Circuit Analysis

Table 23.1 Some f (t), t  0 F(s) ¼ Laplace transform of f(t)


properties of the laplace
kf(t) kF(s)
transform
f1(t) + f2(t) F1(s) + F2(S)
f 0(t) sF(s)  f(0)
00
f (t) s2F(s)  sf(0)  f 0(0)
000 00
f (t) s3F(s)  s2f(0)  sf0(0)  f (0)
Rt
s F ðsÞ
1
f ðxÞdx
0

Recall that for the inductor, the current, and voltage are related by

diL
vL ¼ L :
dt
Its Laplace domain version is (by using Table 23.1 and assuming zero initial
condition)

V L ¼ ðsLÞI L :

This Laplace domain expression is in the form of Ohm’s law with the impedance

Z L ¼ sL,

which is similar to the impedance jω with the phasor notation. In fact, you can just
replace jω with s, you will get the Laplace transform version from the phasor
method’s impedance expression.
Let us do the same for the capacitor, for which the current and voltage are related
by

dvC
iC ¼ C :
dt
Its Laplace domain version is (by using Table 23.1 and assuming zero initial
condition)

I C ¼ ðsC ÞV C :

This Laplsace domain expression is in the form of Ohm’s law with the impedance

1
ZC ¼ ,
sC
1
which is similar to the impedance jωC with the phasor notation. Therefore, by using
the Laplace transform, we can feel freely to use Ohm’s law to set up circuit
equations. Before we can show you a circuit example, we need some Laplace
transform pairs, some of which are listed in Table 23.2.
23 Laplace Transform in Circuit Analysis 173

Table 23.2 Laplace trans- f (t), t  0 F(s) ¼ Laplace transform of f(t)


form pairs
δ(t)23 1
u(t) 1
s
tu(t) 1
s2
eatu(t) 1
sþa
ω
sin(ωt)u(t) s2 þω2
s
cos(ωt)u(t) s2 þω2
teatu(t) 1
ðsþaÞ2
eat sin (ωt)u(t) ω
ðsþaÞ2 þω2
eat cos (ωt)u(t) sþa
ðsþaÞ2 þω2

In Table 23.2, u(t) is the unit step function, defined as (see Fig. 23.1)
8
>
< 0 for t < 0,
uðt Þ ¼ 0:5 for t ¼ 0,
>
:
1 for t > 0:

Fig. 23.1 The unit step function

In Table 23.2, δ(t) is called a unit impulse function, or Dirac delta function, or
simply delta function. The delta function is not really a function in the regular sense.
Anyway, we treat it as a function in this tutorial. It can be defined by the following
two equations:

þ1, t ¼ 0,
δ ðt Þ ¼
0, t 6¼ 0,

and
174 23 Laplace Transform in Circuit Analysis

Z 1
δðt Þdt ¼ 1:
1

It is difficult to graphically show this function. Imagine that ε is a very small


positive number, as we take the limit ε ! 0, the limit of Fig. 23.2 (left) tends to the
delta function. Notice that the area underneath the function curve is always 1. An
official symbol for the delta function is an arrow of length 1 as shown in Fig. 23.2
(right).

Fig. 23.2 The delta function δ(t)

Example
Find the voltage vC(t) after t ¼ 0.

Fig. 23.3 An RC circuit

Solution
After t ¼ 0, a voltage source of 12 V is applied. This is equivalent to a unit step
function with amplitude 12 V. Thus, the Laplace domain version of the circuit in
Fig. 23.3 is shown in Fig. 23.4.
23 Laplace Transform in Circuit Analysis 175

Fig. 23.4 The Laplace


domain version of the circuit
in Fig. 23.3

This is a voltage divider, and thus

1
12 s 12 12 1 3 1
VC ¼ ¼ ¼  ¼  ,
s 20 þ 1s sð1 þ 20sÞ 20 s s þ 20
1 5 s s þ 20
1

! !
3 20 20 1 1
¼ þ ¼ 12  :
5 s s þ 20
1 s s þ 20 1

Looking up Table 23.2 to find the inverse Laplace transform, we have


 t

vC ðt Þ ¼ 12 1  e20 uðt Þ:

Here using u(t) is just another way to say t  0.


20
From s sþ1 1 to get 20s þ sþ 1 is called partial fraction expansion, or partial
ð 20Þ 20

fraction decomposition. There is a short cut to perform it. The reason for doing
partial fraction expansion is that we cannot find s sþ1 1 in Table 23.2; however, both 1s
ð 20Þ
and sþ1 1 are in Table 23.2.
20
To do partial fraction expansion is to find k1 and k2 in

1 k k
 ¼ 1þ 21 :
s s þ 20
1 s s þ 20

To find k1 we multiply s on both sides of the above equation, obtaining

1 sk 2
  ¼ k1 þ :
s þ 20
1
s þ 20 1

Let s ! 0, we have 1
¼ k1 þ 0, that is, k1 ¼ 20.
ð0þ201 Þ 
To find k2 we multiply s þ 20 1
on both sides of the above equation, obtaining
176 23 Laplace Transform in Circuit Analysis

 
1 s þ 20
1
k1
¼ þ k2 :
s s
Let s !  20
1
, we have 1
¼ 0 þ k 2 , that is, k2 ¼  20.
ð201 Þ
Once you are good at it, you can use a cover-up method as explained below. You
start with writing down the equation:

and you want to find k1. On the left-hand side of the equation, you cover up the
denominator, s, of the k1 term as

You realize that as s ! 0 the covered factor will be 0. Pay attention to the left-
hand side, as if the covered part is not there anymore. Let s ! 0 for the rest of the
left-hand side (uncovered part), and the limit is your k1 value 20.  
Next, you want to find k2. You cover up the denominator of the k2 term s þ 201
as

1 k k
 ¼ 1þ 21 :
s s þ 20
1 s s þ 20

You realize that as s !  20


1
the covered factor will be 0. Pay attention to the left-
hand side, as if the covered part is not there anymore. Let s !  201
for the rest of the
left-hand side (uncovered part), and the limit is your k1 value 20.
There are many ways to find the partial fraction expansion (or partial fraction
decomposition). Another way is to sum up the expansion and compare the
coefficients in the numerators on both sides, as shown below.
 
1 k1 k2 k 1 s þ 20
1
þ k 2 s ðk 1 þ k2 Þs þ 201
k1
  ¼ þ ¼   ¼   :
s s þ 20
1 s s þ 20
1
s s þ 201
s s þ 201

By comparing the coefficients on both sides of the equations in the numerators,


we have

1
k1 þ k2 ¼ 0 and k ¼ 1:
20 1
We get the same answer as before: k1 ¼ 20 and k2 ¼ 20.
In fact, the method of comparing coefficients is more useful if the characteristic
polynomial contains complex roots. The characteristic polynomial is simply the
nominator polynomial with respect to s. Let us use one example to see it is done.
23 Laplace Transform in Circuit Analysis 177

Example
Find the inverse Laplace transform of

1
Y ðsÞ ¼ :
sðs2 þ 6s þ 25Þ

Solution
The characteristic polynomial has one real root of s ¼ 0 and two complex conjugate
roots, because
   
s s2 þ 6000s þ 25  106 ¼ s s þ 3Þ2 þ 42 :

In Table 23.2, we have two entries containing the form of (s + a)2 + ω2 in the
denominator. We must use both entries in the expansion.
Thus,

1 k ω ðs þ aÞ
Y ðsÞ ¼ ¼ 1 þ k2 þ k3 ,
sðs2 þ 6s þ 25Þ s 2
ð s þ aÞ þ ω 2 ðs þ aÞ2 þ ω2

k1 4 ð s þ 3Þ
¼ þ k2 þ k3 :
s 2
ð s þ 3Þ þ 4 2
ðs þ 3Þ2 þ 42

The coefficient k1 can be obtained by the cover-up method, and k1 ¼ 1/25. We


have

1
1
4 ð s þ 3Þ
Y ðsÞ ¼ ¼ 25
þ k2 þ k3 :
sðs2 þ 6s þ 25Þ s 2
ð s þ 3Þ þ 4 2
ð s þ 3Þ 2 þ 42

Our next step is to sum up the right-hand side of the above equation, and then
compare the coefficients.

1
1
4 ð s þ 3Þ
¼ 25
þ k2 þ k3 ,
sðs2 þ 6s þ 25Þ s 2
ðs þ 3Þ þ 4 2
ðs þ 3Þ2 þ 42
1
ðs2 þ 6s þ 25Þ þ k2 4s þ k 3 ðs þ 3Þs
¼ 25
:
sðs2 þ 6s þ 25Þ

Numerator coefficients of the s2 term on both sides:


 
1
0¼ þ k3 :
25
Numerator coefficients of the s1 term on both sides:
178 23 Laplace Transform in Circuit Analysis

6
0¼ þ k 2 4 þ k3 3:
25

Numerator coefficients of the s0 term on both sides:

1 ¼ 1:

We obtain

1
k3 ¼  ,
25
and

6
3
3
k2 ¼ 25 25 ¼  :
4 100
According to Table 23.2, the inverse Laplace transform is obtained as

1 3 3t 1
yð t Þ ¼  e sin ð4t Þ  e3t cos ð4t Þ, for t  0:
25 100 25
As another example, if we are asked to find the inverse Laplace transform of

1
Y ðsÞ ¼ ,
s2 ðs2 þ 6s þ 25Þ

its partial fraction exposition is in the form of

1 k k 4 ð s þ 3Þ
Y ðsÞ ¼ ¼ 1 þ 22 þ k3 þ k4 :
s2 ðs2 þ 6s þ 25Þ s s 2
ðs þ 3 Þ þ 4 2
ð s þ 3Þ 2 þ 42
k1
Do not forget the s term because the lower-order terms are required during
expansion.

The inverse Laplace transform of the newer Y(s) is in the form of

yðt Þ ¼ k 1 þ k2 t þ k 3 e3t sin ð4t Þ þ k4 e3t cos ð4t Þ, for t  0:

Oh, wait a minute, we have not told you how the Laplace transform is defined and
how Tables 23.1 and 23.2 are made yet. The Laplace transform of a function f(t) is
defined by an integral as shown in the expression below with a complex variable s.
Z 1
F ðsÞ ¼ f ðt Þest dt:
0

Everything in Tables 23.1 and 23.2 can be verified with this definition.
23 Laplace Transform in Circuit Analysis 179

Notes
Using the Laplace transform, we are able to convert differential equations to
algebraic equations. In the Laplace domain, Ohm’s law works for inductors
and capacitors again. KVL and KCL also work. One can setup the equations
directly in the Laplace domain.
After solving the Laplace-domain algebraic equations, the solution is a
function in the Laplace domain. In order to obtain a time-domain solution
(which a function of time), we need to perform the inverse Laplace transform
on the Laplace-domain solution.
To perform the inverse Laplace transform, the important step is to express
the Laplace-domain function as a summation of the functions in the Laplace
transform pair table (i.e., the right column of Table 23.2). This procedure is
called the partial fraction decomposition, which can be achieved by the cover-
up method or the coefficient comparison method.
Each term in the partial fraction expansion appears in the right column of
Table 23.2. A time-domain expression is then readily obtained by using the
corresponding terms in the left column.
180 23 Laplace Transform in Circuit Analysis

Exercise Problems

Problem 23.1 Solve the following differential equation using the Laplace transform
method.

x} ðt Þ þ 4x0 ðt Þ þ 3xðt Þ ¼ 5

with initial conditions x0(0) ¼ 1 and x(0) ¼ 2.

Problem 23.2 Use the Laplace transform method to solve for the circuit in
Fig. P23.1. The initial voltage in the capacitor is 3 V.

Fig. P23.1

Problem 23.3 Use the Laplace transform method to solve for the circuit in
Fig. P23.2. The initial current in the inductor is 3 A.

Fig. P23.2

Solutions to Exercise problems are given in Book Appendix.


Fourier Transform in Circuit Analysis
24

The Laplace transform is a powerful tool and you may have already seen it
elsewhere, for example, in your Differential Equations class. By using the Laplace
transform, we can do away with the differential equations. However, the Laplace
transform has its drawbacks; it cannot handle the sources that are on the entire time
as in steady-state analysis. By modifying the Laplace transform a little, more general
sources can be accommodated.
Here comes the Fourier transform defined below for function f(t), and we do
NOT have the restriction of t  0.
Z 1
F ð ωÞ ¼ f ðt Þejωt dt:
1

The biggest difference between the Fourier transform and the Laplace transform
is the lower limit of the integral in the definition. If the function f(t) is 0 for t < 0, then
its Fourier transform and its Laplace transform are essentially the same, just
replacing jω by s.
For the Fourier transform, the counterparts of Tables 23.1 and 23.2 are
Tables 24.1 and 24.2, respectively. Notice that the initial conditions no longer appear
in the new tables. It is straightforward to verify that if the function f(t) is an even
function, then its Fourier transform is real and even. If the function f(t) is an odd
function, then its Fourier transform F(ω) is imaginary and odd.
In Table 24.2, sgn(t) is the signum function, which is defined as
8
< 1 for t < 0,
>
sgn ðt Þ ¼ 0 for t ¼ 0,
>
:
1 for t > 0:

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 181
G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_24
182 24 Fourier Transform in Circuit Analysis

Table 24.1 Some f(t) F(ω) ¼ Fourier transform of f(t)


properties of the Fourier
kf(t) kF(ω)
transform
f1(t) + f2(t) F1(ω) + F2(ω)
f 0(t) ( jω)F(ω)
00
f (t) ( jω)2F(ω)
000
f (t) ( jω)3F(ω)
Rt jω F ðωÞ
1
f ðxÞdx
1

Table 24.2 Fourier trans- f(t) F(ω) ¼ Fourier transform of f(t)


form paris
δ(t) 1
1 2πδ(ω)
u(t) πδðωÞ þ jω
1

sgn(t) 2

at
e u(t) 1
sþjω
at
e u(t) 1
sjω
ea j tj 2a
a2 þω2
sin(ω0t) jπ[δ(ω + ω0)  δ(ω  ω0)]
cos(ω0t) jπ[δ(ω + ω0) + δ(ω  ω0)]
e jω0 t 2πδ(ω  ω0)

Example
Use the Fourier transform to find iL(t) in the circuit shown in Fig. 24.1. The
current source is sinusoidal ig(t) ¼ 50 cos (3t) A.

Fig. 24.1 An RL circuit with


a sinusoidal current source for
steady-state analysis
24 Fourier Transform in Circuit Analysis 183

Solution
The Fourier domain equivalent of the circuit in Fig. 24.1 is shown in Fig. 24.2.
According to Table 24.2,

I g ðωÞ ¼ 50π ½δðω þ 3Þ þ δðω  3Þ:

Realizing that the circuit is a current divider, we have

1
I L ðωÞ ¼ I g ðωÞ ,
1 þ ð3 þ jωÞ
1
¼ 50π ½δðω þ 3Þ þ δðω  3Þ ,
4 þ jω

ð4  jωÞ
¼ 50π ½δðω þ 3Þ þ δðω  3Þ ,
ð4 þ jωÞð4  jωÞ
4  jω
¼ 50π ½δðω þ 3Þ þ δðω  3Þ,
16 þ ω2
4  jω 4  jω
¼ 50π δðω þ 3Þ þ 50π δðω  3Þ,
16 þ ω2 16 þ ω2
4  jð3Þ 4  j ð 3Þ
¼ 50π 2
δðω þ 3Þ þ 50π δðω  3Þ,
16 þ ð3Þ 16 þ ð3Þ2

¼ 2π ð4 þ j3Þδðω þ 3Þ þ 2π ð4  j3Þδðω  3Þ,

¼ 8π ½δðω þ 3Þ þ δðω  3Þ þ 6jπ ½δðω þ 3Þ  δðω  3Þ:

Fig. 24.2 An RL circuit with


a sinusoidal current source in
Fourier domain representation
184 24 Fourier Transform in Circuit Analysis

In the above calculation, we employed one unique property that is only valid for
the delta function:

gðxÞδðx  aÞ ¼ gðaÞδðx  aÞ, for any continuous function gðxÞ:

Using Table 24.2 to find the inverse Fourier transform, we have the time domain
expression:

iL ðt Þ ¼ 8 cos ð3t Þ þ 6 sin ð3t Þ A:

In fact, the Fourier transform method is more powerful than the phasor method.
The phasor method can only handle the steady-state sinusoid sources, while the
Fourier transform method can handle any source input that is defined on (1, 1).

Notes
The Fourier transform is almost the same as the Laplace transform. The main
difference is that the Laplace transform requires the time-domain functions
defined in t  0, while the Fourier transform does not have this restriction for
the time-domain functions.
Ohm’s law works for inductors and capacitors by using impedance. In
many cases, we can use the relationship s ¼ jω to change the Laplace-domain
equations to the Fourier-domain equations, and vice versa.
The Laplace transform can easily handle the initial conditions, while the
Fourier transform can analyze steady-state functions.
Exercise Problems 185

Exercise Problems

Problem 24.1 Use the Fourier transform method to solve for the circuit in
Fig. P24.1. The initial voltage in the capacitor is 3 V.

Fig. P24.1

Problem 24.2 The input signal is a signum function. Find the inductor current iL.

Problem 24.3 Repeat Problem 24.2 with 10 cos (4t) being the input signal.

Fig. P24.2

Solutions to Exercise problems are given in Book Appendix.


Second-Order Circuits
25

The voltage and current relationship in a second-order circuit is characterized by a


second-order differential equation. A circuit consisting of a resistor (R), an inductor
(L), and a capacitor (C) is most likely a second order circuit. We have seen an RLC
circuit in action when we discussed steady-state analysis, in which the phasor
notation was used, and the circuit order did not matter.
In this chapter, consider parallel and series RLC circuits with DC sources and
switch actions. In other words, we are interested in the step response of the RLC
circuits. This is not steady-state analysis, and differential equations may need to get
involved. If we try to avoid the differential equations, we can use the Laplace
transform or the Fourier transform.
The step response for a first-order system is fairly easy. It is an exponential
transition from the initial value to the final value, with a time constant to determine
the transition rate. For a second-order system, the transition is more complicated than
an exponential function. The responses are categorized into three types (see
illustrations in Fig. 25.1): the underdamped, critically damped, and overdamped.

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 187
G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_25
188 25 Second-Order Circuits

Fig. 25.1 The step response of a second-order system can have different shapes

Example
Find the step response i(t) of the RLC circuit shown in Fig. 25.2. The initial
conditions are zero. At t ¼ 0, the voltage across the capacitor is zero, and the
current in the inductor is zero.

Fig. 25.2 An RLC circuit

Solution
The initial conditions are zero. In other words, the system is initially relaxed. The
voltage source together with the switch action can be treated as a unit step function.
The Laplace transform of the source is 1/s.
25 Second-Order Circuits 189

We will use the Laplace transform method to solve this problem. We first change
the notation in Fig. 25.2 to Laplace domain notation as shown in Fig. 25.3. Using
Ohm’s law, we have

1 1
s 1 1 L
I ðsÞ ¼ ¼ ¼ :
Rþ 1
sC þ sL s R þ sC
1
þ sL s2 þ RL s þ LC
1

The denominator of I(s) is a second-order polynomial; thus, this is a second-order


circuit. There are three possible outcomes of I(s):

1. [An underdamped solution]

ω sþa
I ðsÞ ¼ k 1 2
þ k2 ,
ð s þ aÞ þ ω 2 ðs þ aÞ2 þ ω2

which corresponds to

iðt Þ ¼ k 1 eat sin ðωt Þuðt Þ þ k 2 eat cos ðωt Þuðt Þ:

Since

1
L
I ðsÞ ¼ R
s2 þ Ls þ LC
1

and

ω sþa
I ðsÞ ¼ k 1 2
þ k2 ,
ð s þ aÞ þ ω 2 ðs þ aÞ2 þ ω2

their nominators must be identical, that is,

1
¼ k 1 ω þ k 2 ðs þ aÞ:
L
By comparing the coefficients, we must have

1
k2 ¼ 0 and k1 ¼ :

Thus,

iðt Þ ¼ k 1 eat sin ðωt Þuðt Þ

as shown in Fig. 25.4. In order to have an underdamped solution, the character-


istic equation (setting the denominator to zero) s2 þ RL s þ LC
1
¼ 0 must have two
complex conjugate solutions: s1, 2 ¼  a  jω. That is, we must have
190 25 Second-Order Circuits

 2
R 4
 <0
L LC
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 R 2
R
with a ¼ 2L and ω ¼ LC 1
 2L .
2. [A critically damped solution]

1 1 k ð s þ aÞ þ k 4
I ðsÞ ¼ k 3 þ k4 ¼ 3 ,
sþa ð s þ aÞ 2
ð s þ aÞ 2

which corresponds to

iðt Þ ¼ k 3 eat uðt Þ þ k 4 teat uðt Þ:

Since

1
L
I ðsÞ ¼ R
s2 þ Ls þ LC
1

and

k 3 ð s þ aÞ þ k 4
I ðsÞ ¼ ,
ð s þ aÞ 2

their nominators must be identical, that is,

1
¼ k 3 ð s þ aÞ þ k 4 :
L
By comparing the coefficients, we must have

1
k3 ¼ 0 and k4 ¼ :
L
Thus,

1
iðt Þ ¼ teat uðt Þ
L
as shown in Fig. 25.4. In order to have a critically damped solution, the charac-
teristic equation s2 þ RL s þ LC
1
¼ 0 must have two identical real solutions: s1,
2 ¼  a. That is, we must have
 2
R 4
 ¼0
L LC
with
25 Second-Order Circuits 191

R 1
a¼ ¼ pffiffiffiffiffiffi :
2L LC
3. [An over-damped solution]

1 1 k ð s þ bÞ þ k 6 ð s þ aÞ
I ðsÞ ¼ k5 þ k6 ¼ 5 ,
sþa sþb ð s þ aÞ ð s þ bÞ

which corresponds to

iðt Þ ¼ k5 eat uðt Þ þ k6 ebt uðt Þ:

Since

1
L
I ðsÞ ¼ R
s2 þ Ls þ LC
1

and

k 5 ð s þ bÞ þ k 6 ð s þ aÞ
I ðsÞ ¼ ,
ðs þ aÞðs þ bÞ

their nominators must be identical, that is,

1
¼ k 5 ðs þ bÞ þ k 6 ðs þ aÞ:
L
By comparing the coefficients, we must have

1
k5 ¼ k6 and k 5 b þ k6 a ¼ :
L
They lead to

1
k 5 ð b  aÞ ¼ ,
L
1
k5 ¼ ,
Lðb  aÞ
1
k6 ¼ :
Lða  bÞ

Thus,

1  at 
iðt Þ ¼ e  ebt uðt Þ
Lðb  aÞ
192 25 Second-Order Circuits

as shown in Fig. 25.4. In order to have an over-damped solution, the characteristic


equation s2 þ RL s þ LC
1
¼ 0 must have two different real solutions: s1 ¼  a and
s2 ¼  b. That is, we must have
 2
R 4
 > 0,
L LC
R 1
with a þ b ¼ and ab ¼
L LC
or
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi
R R 2 R R 2
L þ L  4
LC L  L  4
LC
a¼ and b ¼ :
2 2

Fig. 25.3 An RLC circuit


using Laplace transform
notation

Fig. 25.4 Three potential outcomes of the current in the circuit of Fig. 25.2 or Fig. 25.5
25 Second-Order Circuits 193

Finally, we consider a natural response of a circuit. A natural response is the


response due to the initial conditions. There are no external sources.

Example
Consider the RLC circuit shown in Fig. 25.5. The initial conditions are i
(0) ¼ 0.5 A and v(0) ¼ 0 V. Find the natural response v(t) for t  0.

Fig. 25.5 An RLC circuit


with initial current i(0) ¼ 0.5
A and initial voltage v
(0) ¼ 0 V

Solution
diðt Þ
In Table 1 we learned that the Laplace transform of dt is sI(s)  i(0). For the
inductor, we have vðt Þ ¼ L didtðtÞ.
Thus, in the Laplace domain, this relationship is V
(s) ¼ LsI(s)  Li(0). This implies that if there is a non-zero initial condition, we must
add a source Li(0) to the circuit. The relationship V(s) ¼ LsI(s)  Li(0) also implies
that Li(0) is an independent voltage source. Therefore, we can obtain a Laplace
domain equivalent circuit as in Fig. 25.6.
If we compare the voltage sources in Figs. 25.3 and 25.6, they are different. The
source in Fig. 25.3 is 1/s and in the time domain it is a step function according to
Table 2 in Chap. 23. A step function power source keeps providing energy to the
circuit once it is on. On the other hand, the source in Fig. 25.6 is Li(0) (without the
1/s part) and in the time domain it is a delta function according to Table 2 in
Chap. 23. A delta function power source only provides energy to the circuit at one
moment and provides no energy to the circuit after that.
When using the Laplace transform approach, the initial conditions are converted
into dependent sources, and then the circuits are assumed to have zero initial
conditions.
194 25 Second-Order Circuits

Fig. 25.6 Laplace domain


equivalent circuit for the
circuit in Fig. 25.5

We can now set up a node-voltage equation as

V ðsÞ V ðsÞ þ Lið0Þ


þ sCV ðsÞ þ ¼ 0,
R sL
which leads to

ið0Þ
s i ð 0Þ i ð 0Þ 1
V ðsÞ ¼ ¼ ¼ :
1
R þ sC þ 1
sL Rs
1
þ Cs þ L
2 1 C s þ RC s þ LC
2 1 1

The characteristic equation is

1 1
s2 þ sþ ¼ 0,
RC LC
which can have three different types of solutions depending on the values of the R, L,
and C.

1. Two complex conjugate solutions, when ðRC1 Þ2 < LC


4
. [Underdamped]
The solution is in the form of

vðt Þ ¼ k1 eat sin ðωt Þuðt Þ þ k2 eat cos ðωt Þuðt Þ:

Since

i ð 0Þ 1
V ðsÞ ¼
C s2 þ RC
1
s þ LC
1

and
25 Second-Order Circuits 195

ω sþa
V ðsÞ ¼ k 1 þ k2 ,
ð s þ a Þ 2 þ ω2 ð s þ a Þ 2 þ ω2

their nominators must be identical, that is,

ið0Þ
¼ k1 ω þ k 2 ðs þ aÞ:
C
By comparing the coefficients, we must have

i ð 0Þ
k 2 ¼ 0 and k 1 ¼ :

Thus,

ið0Þ at
vð t Þ ¼ e sin ðωt Þuðt Þ

as shown in Figure 25.4.
2. Two identical real solutions, when 1
ðRC Þ2
¼ LC
4
. [Critically damped]
The solution is in the form of

vðt Þ ¼ k3 eat uðt Þ þ k4 teat uðt Þ:

Since

i ð 0Þ 1
V ðsÞ ¼
C s2 þ RC
1
s þ LC
1

and

1 1 k ð s þ aÞ þ k 4
V ðsÞ ¼ k3 þ k4 ¼ 3 ,
sþa ð s þ aÞ 2
ðs þ aÞ2

their nominators must be identical, that is,

i ð 0Þ
¼ k 3 ð s þ aÞ þ k 4 :
C
By comparing the coefficients, we must have

i ð 0Þ
k 3 ¼ 0 and k 4 ¼ :
C
Thus,
196 25 Second-Order Circuits

ið0Þ at
vð t Þ ¼ te uðt Þ
C
as shown in Fig. 25.4.
3. Two different real solutions, when ðRC1 Þ2 > LC
4
. [Over-damped]
The solution is in the form of

vðt Þ ¼ k5 eat uðt Þ þ k6 ebt uðt Þ:

Since

i ð 0Þ 1
V ðsÞ ¼
C s2 þ RC
1
s þ LC
1

and

k 5 ð s þ bÞ þ k 6 ð s þ aÞ
V ðsÞ ¼ ,
ðs þ aÞðs þ bÞ

their nominators must be identical, that is,

ið0Þ
¼ k 5 ðs þ bÞ þ k 6 ðs þ aÞ:
C
By comparing the coefficients, we must have

ið0Þ
k5 ¼ k 6 and k 5 b þ k6 a ¼ :
C
They lead to

ið0Þ
k 5 ð b  aÞ ¼ ,
C
i ð 0Þ
k5 ¼ ,
C ð b  aÞ

ið0Þ
k6 ¼ :
C ð a  bÞ

Thus,

ið0Þ  at 
vð t Þ ¼ e  ebt uðt Þ
C ð b  aÞ

as shown in Fig. 25.4.


25 Second-Order Circuits 197

You may have noticed that sometimes we use the notation f(0) and other times we
use the notation f(0), for example, in Table 1 of Chap. 23. Are they the same? If the
function f(t) is not allowed to have a sudden change of values such as the inductor
current or capacitor voltage, we have f(0) ¼ f(0) ¼ f(0+), and we can use either f(0)
or f(0) to indicate the initial condition. In general, use f(0) for the initial condition if
you are not sure whether f(0) and f(0+) are the same.

Notes
A typical second-order circuit consists of RLC and is described by a second-
order differential equation. The response to a simple switch action can be
underdamped, critically damped, or over-damped, depending on the values of
R, L, and C.
The superposition principle allows us to study the natural response and step
response. For the natural response, the initial values are treated as delta
function power sources. In the step response studies, the initial conditions
are assumed to be zero.
The Laplace transform is effective in these transient response studies.
198 25 Second-Order Circuits

Exercise Problems

Problem 25.1 In a series RLC circuit, what does the step response look like when
R ¼ 0? The input is step voltage source, and the output is the voltage across the
capacitor.

Fig. P25.1

Problem 25.2 In a parallel RLC circuit, what does the step response look like when
R ¼ 1? The input is step current source, and the output is the current through the
inductor.

Fig. P25.2
Exercise Problems 199

Problem 25.3 The circuit in Fig. P25.3 is neither a series nor a parallel circuit. It is
still a second-order RLC circuit. Find the conditions for circuit to be underdamped,
critically damped and over-damped.

Fig. P25.3

Solutions to Exercise problems are given in Book Appendix.


Filters
26

This chapter revisits sinusoidal steady-state analysis, emphasizing the frequency


dependency. In Chap. 19, the emphasis was on a fixed given frequency; the
frequency stays the same throughout the entire circuit.
This chapter assumes that the circuit has only one steady-state sinusoidal voltage
source and its frequency can be anywhere from 0 to 1. The output of the circuit is a
voltage. We are interested in the ratio of the output voltage over the input voltage as a
function of the frequency ω. We are more interested in the magnitude (i.e., the norm)
of the ratio, which is referred to as the gain. The gain is the same, whether the circuit
is represented in time domain, Fourier domain, or phasor notation.
We can either use the phasor method or the Fourier transform method to find the
output voltage. Actually, these two approaches are the same for filter design and
analysis. To get started, let us consider a simple RC circuit in Fig. 26.1. This is a
voltage divider, and the gain is readily expressed as
   1 
V out ðωÞ  jωC  1 1

gainðωÞ ¼  ¼ ¼ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi :
V in ðωÞ  R þ jωC
1  jjωRC þ 1 j
ðωRC Þ2 þ 1

Let us look at two extreme cases of ω ¼ 0 and ω ! 1, respectively. When ω ¼ 0,


the gain reaches its maximum value of 1. When ω ! 1, the gain approaches its
maximum value of 0. The circuit acts like a lowpass filter, having a higher gain at
low frequencies and lower gain at high frequencies (see Fig. 26.2).

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 201
G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_26
202 26 Filters

Fig. 26.1 An RC circuit


represented in the Fourier
domain

Fig. 26.2 Frequency


response of a low-pass filter

If we switch the positions of the capacitor and the resistor in Fig. 26.1, the new
circuit (see Fig. 26.3) will behave like a highpass filter (see Fig. 26.4), in the sense
that the gain is higher at high frequencies and the gain is lower at low frequencies. In
fact, the circuit in Fig. 26.3 is voltage divider, and its gain is evaluated as

Fig. 26.3 An RC circuit


represented in the Fourier
domain
26 Filters 203

Fig. 26.4 Frequency


response of a high-pass filter

   
V out ðωÞ  R  jjωRC j ωRC

gainðωÞ ¼   ¼ ¼ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi :
V in ðωÞ  R þ jωC jjωRC þ 1j
1
ðωRC Þ2 þ 1

When ω ¼ 0, the gain reaches its minimum value of 0. When ω ! 1, the gain
approaches its maximum value of 1.
Circuits that specifically attenuate some frequencies and enhance other
frequencies are called filters. Filters consisting of resistors, capacitors, and inductors
are passive filters because they do not require an external power source (beyond the
input signal). Active filters, on the other hand, require external power source to
operate. Most active filters use op-amps. An inverting amplifier is shown in
Fig. 26.5, and its gain is given as
   
v   R f  R f
gain ¼  out  ¼   ¼ :
vin Ri Ri

Fig. 26.5 An inverting amplifier using resistors


204 26 Filters

Fig. 26.6 A general inverting amplifier

If we replace the resistors in Fig. 26.5 by general impedances as shown in


Fig. 26.6 with the gain
   
V   Z f 
gain ¼  out  ¼  ,
V in Zi

we can obtain many useful active filters. One special case is shown in Fig. 26.7, and
its gain is

Fig. 26.7 A bandpass filter


26 Filters 205

Fig. 26.8 Frequency


response of a bandpass filter

     
V out   R
 
  R

R
gain ¼   ¼ 
  ¼    ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 ffi:
V in  R þ jωL þ jωC  R þ j ωL  ωC 
1 1
1 2
R þ ωL  ωC
2

When ωL ¼ ωC 1
or ω ¼ p1ffiffiffiffi
LC
ffi, the gain reaches to its maximum value of 1. When
ω ! 0 or ω ! 1, the gain tends to its minimum value of 0. Therefore, this is a
bandpass filter (see Fig. 26.4). The frequency ω0 ¼ p1ffiffiffiffi LC
ffi is called the resonance
radian frequency, at which the effects of a capacitor and an inductor cancel out
(they can be replaced by a short circuit). At the resonance frequency, the circuits in
Figs. 26.1 and 26.3 are the same with Rf ¼ Ri ¼ R (Fig. 26.8).
Likewise, the circuit shown in Fig. 26.9 is another bandpass filter configuration
with the gain calculated as
   
   
V out   R1þjωL1 þjωC  
1
1  1
gain ¼   ¼   ¼   
1 
¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 :
V in   R   1 þ jR ωC  ωL 
1 þ R2 ωC  1 ωL

At the frequency ω0 ¼ p1ffiffiffiffi


LC
ffi, the effects of a capacitor and an inductor cancel out
(they can be replaced by an open circuit).
Finally, we would like to mention that people usually express the filter gain in
“dB” (decibel), which is defined by 20 times the common logarithm (that is, the
logarithm to base 10) as
 
v 
Number of decibels ¼ 20 log 10  out :
vin
206 26 Filters

Fig. 26.9 A bandpass filter

For example, a voltage gain of 1 is 0 dB; a voltage gain of 10 is 20 dB; a voltage


gain of 100 is 40 dB. If the voltage gain is less than 1, the dB value is negative. For
example, the voltage gain of 0.1 corresponds to 20 dB.

Notes
In this chapter, the focus is the circuit gain as a function of the frequency. A
circuit that intentionally enhances some frequencies and attenuates other
frequencies is called a filter. To study the filter behavior is in the category of
sinusoidal steady-state analysis. The Fourier transform (or equivalently, the
phasor transform) is a natural fit for this type of studies.
It is interesting to notice that for a second-order RLC filter, there is a certain
frequency, at which the effects of the capacitor and inductor cancel each other
out. This frequency is called the resonant frequency of the filter.
Exercise Problems 207

Exercise Problems

Problem 26.1 Without doing any mathematical derivation, determine whether the
Sallen-Key filter shown in Fig. P26.1 a lowpass filter or a highpass filter.

Fig. P26.1
208 26 Filters

Problem 26.2 Without doing any mathematical derivation, determine whether the
Sallen-Key filter shown in Fig. P26.2 a lowpass filter or a highpass filter.

Fig. P26.2

Problem 26.3 Without doing any mathematical derivation, determine whether the
Sallen-Key filter shown in Fig. P26.3 a lowpass filter or a highpass filter or none of
them.

Fig. P26.3

Solutions to Exercise problems are given in Book Appendix.


Wrapping Up
27

Ohm’s law is the foundation of electrical engineering. The resistors obey Ohm’s law.
We can use KCL and KVL to set up linear equations and solve for the unknowns.
Circuits can be simplified by using Thévenin equivalent and Norton equivalent
circuits. If there is a dependent source in the circuit, we always need an equation
to describe that dependent source. An op-amp is a dependent source. An ideal
op-amp is a good approximation of a practical op-amp. For a stable op amp circuit,
the feedback path should go to the inverting input labeled by a “” sign.
You do not have to solve a circuit problem by solving system of linear equations.
Many times, a simple voltage divider or current divider method can give you the
answer quickly.
Everything goes well until capacitors and inductors show up. The voltage and
current for a capacitor or an inductor are related with a derivative or an integral, and
Ohm’s law is not obeyed. The KCL and KVL equations become differential
equations. We admit that it is not easy to work with differential equations. This is
where the various transform methods come in. The sole purpose of these methods is
to make Ohm’s law work again for capacitors and inductors so that the differential
equations reduce to algebraic equations.
The Laplace transform can be used when we are considering initial conditions
and switch actions. In the Laplace domain, capacitors follow Ohm’s Law V C ¼
sC I C and inductors follow Ohm’s Law VL ¼ (sL)IL. The Fourier transform is useful
1

for steady-state analysis and frequency response studies. If the source is pure
sinusoidal, the phasor method is a method dedicated to evaluating the magnitude
response and phase shift. In the Fourier domain and the phasor method, capacitors
follow Ohm’s Law V C ¼ jωC 1
I C and inductors follow Ohm’s Law VL ¼ ( jωL)IL.
Basically, you either use “s” or use “jω” for the time-domain derivative. Filter
analysis is easier to perform by using “jω.”
Every periodic function can be expanded into a Fourier series. Each term in the
expansion is sinusoidal, and thus the phasor method can be applied to each term in
the Fourier series. The superposition principle can be used to obtain the result if the

# The Author(s), under exclusive license to Springer Nature Switzerland AG 2021 209
G. L. Zeng, M. Zeng, Electric Circuits,
https://doi.org/10.1007/978-3-030-60515-5_27
210 27 Wrapping Up

input is a general periodic function. In fact, the Fourier transform and the Fourier
series are closely related. The Fourier transform is defined for non-periodic
functions, and the Fourier series for periodic functions. The Fourier transform is
the limiting case for the Fourier series by letting the period T ! 1.
For a first-order circuit, you can just write down the time-domain step response of
a capacitor voltage or an inductor current without setting up equations. The voltage
across a capacitor and the current in an inductor cannot have a sudden jump. They
can change slowly from the initial value to the final value along an exponential path,
governing by a time constant.
The time-domain step response of a second-order circuit also takes time from the
initial value to the final value but following one of the three types of paths. One type
is called the “underdamped” in which the response curve oscillates up-and-down
many times before settling down as combined eat cos (ωt) and eat sin (ωt)
functions. Another type is called the “overdamped” in which the response curve
has the least or no oscillations at all; it is a combination of two exponential functions
eat and ebt functions. The third type is called the “critically damped,” and the
response is a combination of an eat curve and a teat curve.
Any circuit can be a filter. A way to describe a filter is by using a gain function.
The gain is the magnitude of the transfer function. The transfer function is the ratio
of the output over the input in the Fourier or Laplace domain.

Notes
Circuit analysis requires solving KVL and KCL equations. When the trans-
form methods are used, Ohm’s law works for inductors, capacitors, and
resistors. The KVL and KCL equations are now algebraic equations. The
transform-domain solution needs to be converted back to the time domain.
This step is usually achieved by the partial fraction decomposition and by table
lookup.
Exercise Problems 211

Exercise Problems

Problem 27.1 Are KVL equations and mesh equations the same? Are KCL
equations and node equations the same.

Problem 27.2 Do you have a preference regarding to nodes equations or mesh


equations?

Problem 27.3 The main purpose of the Laplace transform and the Fourier transform
is to avoid solving differential equations in the time domain. Why do we need both
the Laplace transform and the Fourier transform? Can we just learn one of them?

Problem 27.4 Which is more powerful, the Fourier transform method or the phasor
method?

Problem 27.5 Name one most important concept in electric circuits.

Solutions to Exercise problems are given in Book Appendix.


Appendix. Solutions to Exercise Problems

Chapter 1. Voltage, Current, and Resistance

Problem 1.1 Either of the following two symbols represents a DC voltage source.
Here “V” is an abbreviation of “Volts.” “Volt” is a unit of voltage.

Fig. P1.1a

Determine whether the following configurations of voltage sources are valid or


invalid. Why?

Solution
(a) Valid. Two (or more) DC voltage sources can be connected in parallel. There is
no current flowing in the circuit.
(b) Invalid. These two voltage sources with different voltages cannot be connected
in parallel.
(c) Valid. Voltage sources can be connected in series. The total voltage is the sum of
each voltage source. Vtotal in this case is 4 V.
(d) Valid. Voltage sources can be connected in series, even though the source
polarities are not consistent. The opposite polarity results in a negative voltage.
The total voltage in this case is 2 V.

# The Editor(s) (if applicable) and The Author(s), under exclusive license to 213
Springer Nature Switzerland AG 2021
G. L. Zeng, M. Zeng, Electric Circuits, https://doi.org/10.1007/978-3-030-60515-5
214 Appendix. Solutions to Exercise Problems

Fig. P1.1b

Problem 1.2 The purpose of a voltage source in a circuit is to cause the current to
flow in a circuit. The flow of the electric current can be converted into something
useful to us. For example, the electric current running through a heating wire can
generate heat. The electric current running through a light bulb creates light. The
electric current running through an electric motor causes motion. Please comment on
the circuit shown whether this circuit is useful.

Fig. P1.2
Appendix. Solutions to Exercise Problems 215

Solution
This circuit should never be allowed because voltage source is shorted. One should
never ever short a voltage source!
This situation is similar to Case (b) in Problem 1.1, the above circuit is equivalent
to two voltage sources connected in parallel, one with a voltage of 3 V and the other
with a voltage of 0 V.

Fig. S1.2

The voltage sources in the textbooks are ideal, in the sense that their voltages keep
constant regardless the rest of the circuit. In our everyday life, batteries and power
supplies are close approximations of the ideal voltage sources. DO NOT short them!
Otherwise, the fuse of the power supplies may be blown. The house circuit breaker
may be trigged. Fire may be caused.

Problem 1.3 Even though we do not see them in everyday life, there are such things
called “current sources.” The ideal current source provides constant current, regard-
less the rest of the circuit. The symbol for a current source is shown below. Here “A”
is an abbreviation of “Amperes.” “Amperes” is a unit of current.

Fig. P1.3a
216 Appendix. Solutions to Exercise Problems

Determine whether the following configurations of current sources are valid or


invalid. Why?

Fig. P1.3b

Solution
(a) Valid. Two or more current sources can be connected in series as long as their
current values are the same.
(b) Invalid. If the current values are not the same, they cannot be connected in
series. They provide conflicting current values.
(c) Valid. Current sources can be connected in parallel. Since these two current
sources are connected with opposite polarity, the total current to the difference
of the currents provided by the two current sources. This this case, the total
current is 3 A.
(d) Valid. The total current is the sum of the two sources when their polarities are
the same. In this case, the total current is 5 A.
Appendix. Solutions to Exercise Problems 217

Problem 1.4 Determine whether the following circuits are valid.

Fig. P1.4

Solution
(a) Valid. This circuit is equivalent to the circuit shown below. This circuit is
different from that in Problem 1.1(b), in which there is no resistor in the circuit.
The resistor here is labeled with “10 Ω.” The unit of resistance is Ohm (Ω).

Fig. S1.4

(b) Valid. The voltage source maintains its voltage regardless the current drawn
from it and it allows be connected to a current source. Likewise, the current
source maintains is current output regardless the voltage.

Problem 1.5 Draw a schematic for the flashlight circuit.

Solution
The schematic for the flashlight is shown below. The light bulb acts as a resistor.

Fig. S1.5
218 Appendix. Solutions to Exercise Problems

Chapter 2. DC Power Supply and Multimeters

Problem 2.1 You are given a power supply and a circuit schematic shown. Suggest
three ways to connect the power supply to the 1 kΩ resistor.

Fig. P2.1

Solution
The three ways are illustrated below:
(a)

Fig. S2.1a

(b)

Fig. S2.1b
Appendix. Solutions to Exercise Problems 219

(c)

Fig. S2.1c

Problem 2.2 Identify the mistakes in using a multimeter.

(a) Trying to measure the voltage across the power supply.

Fig. P2.2a
220 Appendix. Solutions to Exercise Problems

(b) Trying to measure the current through the resistor.

Fig. P2.2b

(c) Trying to measure the resistance of the resistor.

Fig. P2.2c

Solution
(a) In a power supply, the GND (ground) terminal is the chassis ground terminal.
Depending on the application, this chassis ground may be connected to the
negative () terminal, may be connected to the positive (+) terminal, or may not
be connected to any terminals (floating). The DC power output is through the
positive (+) and negative () terminals. A correct connection to measure the
output voltage is shown in the following figure.
Appendix. Solutions to Exercise Problems 221

Fig. S2.2a

(b) Never ever do this when measure current. This will burn the multimeter or its
fuse. The ammeter has very low resistance. When an ammeter is directly
connected to the power supply output, a huge current will run through the
meter. In order to measure the current through a resistor, the ammeter must be
connected in series with the resistor as shown below.

Fig. S2.2b

(c) When measuring the resistance of a resistor, we must remove the resistor from
the circuit and measure the resistor by itself as shown below for two reasons. The
first reason is that the resistor may be connected to other components in the
circuit is a rather complicated way. For example, this resistor may be connected
to a smaller resistor in parallel. The reading of the measurement will be affected
by other components connected to your resistor of interest. The second reason is
that in the resistance measuring mode, the ohm meter is actually measuring
current while using meter’s internal power to supply the voltage. If the circuit’s
222 Appendix. Solutions to Exercise Problems

power is on, the current running through the resistor will affect the resistance
reading.

Fig. S2.2c

Chapter 3. Ohm’s Law

Problem 3.1 Use Ohm’s law to calculate the current in the circuit.

Solution
(a) The voltage source polarity is consistent with the current direction, and thus

V 10 V
I¼ ¼ ¼ 2 A:
R 5Ω
The actual current is running from left-to-right through the 5 Ω resistor.
(b) The voltage source polarity is consistent with the current direction, but the
voltage source has a negative value. We thus have

V 10 V
I¼ ¼ ¼ 2 A:
R 5Ω
Therefore, the actual current is running from right-to-left through the 5 Ω
resistor. The negative sign in the answer “2 A” implies that the current is
running in the opposite direction as that indicated by the arrow in the figure.
(c) The voltage source polarity is inconsistent with the current direction, and thus
we have

V ð10 VÞ
I¼ ¼ ¼ 2 A:
R 5Ω
Appendix. Solutions to Exercise Problems 223

Fig. P3.1

The negative sign in the answer “2 A” implies that the current is running in the
opposite direction as that indicated by the arrow in the figure.
(d) The voltage source polarity is inconsistent with the current direction, and the
voltage value of the voltage source is negative. We now have a case of “double
negative.” Thus, we have

V ð10 VÞ
I¼ ¼ ¼ 2 A:
R 5Ω
224 Appendix. Solutions to Exercise Problems

The positive answer “2 A” implies that the current is running in the direction as
that indicated by the arrow in the figure.
(e) The voltage source polarity is inconsistent with the current direction, and thus
we have

V ð10 VÞ
I¼ ¼ ¼ 2 A:
R 5Ω
The negative sign in the answer “2 A” implies that the current is running in the
opposite direction as that indicated by the arrow in the figure. The actual current
is running from left-to-right through the 5 Ω resistor.
(f) The voltage source polarity is inconsistent with the current direction, and the
voltage value of the voltage source is negative. We now have another case of
“double negative.” Thus, we have

V ð10 VÞ
I¼ ¼ ¼ 2 A:
R 5Ω
The positive answer “2 A” implies that the current is running in the direction as
that indicated by the arrow in the figure. The current is running from right-to-left
through the 5 Ω resistor.
(g) The voltage source polarity is consistent with the current direction, and thus

V 10 V
I¼ ¼ ¼ 2 A:
R 5Ω
The actual current is running from right-to-left through the 5 Ω resistor.
(h) The voltage source polarity is consistent with the current direction, but the
voltage source value is negative. Thus,

V 10 V
I¼ ¼ ¼ 2 A:
R 5Ω
The actual current is running from left-to-right through the 5 Ω resistor.

Problem 3.2 According to the partial circuit shown, use Ohm’s law to calculate the
voltage across the resistor. You must use the voltage polarity and current direction
specified in the figure.

Solution
Ohm’s law can only be applied to a resistor. The “voltage V” in Ohm’s law is always
the voltage across the resistor. The “current I” in Ohm’s law is always the current
through the resistor. The relationship V ¼ IR holds when the voltage polarity and the
current direction are consistent: The current arrow points from “+” to “” of the
voltage labels on the two ends of the resistor.
Appendix. Solutions to Exercise Problems 225

Fig. P3.2

(a) The notation is consistent, and the current is positive. Thus,

V ¼ IR ¼ ð3 AÞð2 ΩÞ ¼ 6 V:
(b) The notation is inconsistent, and a negative sign is required to compensate for
this inconsistency. Thus,
226 Appendix. Solutions to Exercise Problems

V ¼ IR ¼ ð3 AÞð2 ΩÞ ¼ 6 V:

The voltage is negative, which means that the right end of the resistor has a
higher electric potential than the potential at the left end of the resistor.
(c) The notation is consistent, but the current is negative. We have

V ¼ IR ¼ ð3 AÞð2 ΩÞ ¼ 6 V:

The voltage is negative, which means that the right end of the resistor has a
higher electric potential than the potential at the left end of the resistor. The
current is actually running from right-to-left through the 2 Ω resistor.
(d) The notation is inconsistent, and the current is negative. We have a “double-
negative” case,

V ¼ IR ¼ ð3 AÞð2 ΩÞ ¼ 6 V:

The voltage is positive, which means that the left end of the resistor has a higher
electric potential than the potential at the right end of the resistor.
(e) The notation is inconsistent, and a negative sign is required for Ohm’s law. We
have

V ¼ IR ¼ ð3 AÞð2 ΩÞ ¼ 6 V:

The voltage is higher on the left. The actual current always runs from high to low
in a resistor.
(f) The notation is consistent. We have

V ¼ IR ¼ ð3 AÞð2ΩÞ ¼ 6 V:
(g) The notation is inconsistent, and the current is negative. We have

V ¼ IR ¼ ð3 AÞð2ΩÞ ¼ 6 V:


(h) The notation is consistent, but the current is negative. Thus,

V ¼ IR ¼ ð3 AÞð2 ΩÞ ¼ 6 V:

Problem 3.3 You are given an electrical element without any labels. You connect
the element with a variable voltage source. You make some voltage/current
measurements as shown in the table. What most likely is this element?

V (Volts) I (Amperes)
10 5
0 0
10 5
Appendix. Solutions to Exercise Problems 227

Fig. P3.3

Solution
We first plot the voltage vs current curve, which looks like a straight line.

Fig. S3.3

The slope of this straight is R ¼ 2 V/A. The measurements consistently follow


Ohm’s law V ¼ IR, with R ¼ 2 Ω. Hence, this element behaves like a 2 Ω resistor.

Problem 3.4 True or False?

(a) If you double the voltage across the resistor, the current through it doubles.
(b) If you double the voltage across the resistor, the current through it halves.
(c) If you halve the current through the resistor, the voltage across it doubles.
(d) If you halve the current through the resistor, the voltage across it halves.
(e) If you double the resistance of a resistor and keep the voltage across the resistor
unchanged, the current through the resistor doubles.
(f) If you double the resistance of a resistor and keep the current through the resistor
unchanged, the voltage across the resistor doubles.
228 Appendix. Solutions to Exercise Problems

Solution
(a) True.
(b) False. The current doubles.
(c) False. The voltage halves.
(d) True
(e) False. The current halves.
(f) True.

Problem 3.5 The total human body in water is approximately 300 Ω. The electric
current over 10 mA is life threatening if the current runs through the heart
(10 mA ¼ 0.01 A). How much voltage in the water can be lethal?

Solution
According to Ohm’s law,

V ¼ IR ¼ ð10 mAÞð300 ΩÞ ¼ 3000 mV ¼ 3 V:

This lethal voltage of 3 V is very low.


The good news is that dry skin has resistance as high as 100 kΩ (1 kΩ ¼ 1000 Ω).
Dry and unbroken skin can protect us at low voltage. But high voltage (above 500 V)
can breakdown our skin.

Chapter 4. Kirchhoff’s Voltage Law (KVL)

Problem 4.1 Using the given voltage polarities, set up KVL equations for the
following circuits:

Fig. P4.1
Appendix. Solutions to Exercise Problems 229

Solution
There are two meshes in each circuit, and we can set up two KVL equations for each
case.

(a) v0 + v2 ¼ 3 V,
v1 þ v2 ¼ 1 V:

(b) v0  v2 ¼ 3 V,
v1  v2 ¼ 1 V:

(c) v0  v2 ¼ 3 V,
v1  v2 ¼ 1 V:

(d) v0 + v2 ¼  3 V,
v1 þ v2 ¼ 1 V:

Problem 4.2 In this problem, we will use a new current source called controlled
current source, whose symbol is a diamond with an arrow inside (see the figure
below). The symbol for a regular current source is a circle with an arrow inside. For
example, a controlled current source is

Fig. P4.2a

Here “2i” indicate the value of this current source, and this value is two times the
current value i, which is defined elsewhere in the circuit.

Fig. P4.2b

Find the current i in this circuit.


230 Appendix. Solutions to Exercise Problems

Solution
Since the elements in series with a current source have no contribution to the circuit,
we can ignore the 3 Ω resistor and the 1 V voltage source. [Likewise, we can ignore
the elements in parallel with a voltage source.]
The current running down the 2 Ω resistor is 3i, which is the summation of i from
the left branch and 2i from the right branch.
Using Ohm’s law, the voltage across the 1 W resistor is (1 Ω)(i).
The voltage across the 2 W resistor is (2 Ω)(3i).
The KVL equation for the left mesh is given as

ð1 ΩÞðiÞ þ ð2 ΩÞð3iÞ ¼ 2 V:

That is,

7i ¼ 2 V:

Finally,

i ¼ 2=7 A:

Problem 4.3 Set up the KVL equations for the following Wheatstone bridge circuit.

Fig. P4.3

Solution
This circuit consists of three meshes, and we can set up three KVL equations.

v1 þ v3 ¼ 10 V,

v1 þ v5  v2 ¼ 0,

v5 þ v4  v3 ¼ 0:
Appendix. Solutions to Exercise Problems 231

Problem 4.4 Do not simplify the circuit. Use the KVL to solve for the current i in
the circuit.

Fig. P4.4

Solution
Using Ohm’s law, the voltage drop across the 1 Ω resistor is (1 Ω)(i).
The voltage drop across the 3 Ω resistor is (3 Ω)(i).
The circuit has only one mesh, and we can set up one KVL equation as follows:

10 V ¼ ð1 ΩÞðiÞ þ ð2 VÞ þ ð3 ΩÞðiÞ:

After simplification, we obtain

8 V ¼ ð4 ΩÞðiÞ,

and

i ¼ 2 V=Ω ¼ 2 A:
232 Appendix. Solutions to Exercise Problems

Problem 4.5 Use KVL to verify if the following circuit is valid.

Fig. P4.5

Solution
If an element is connected to a voltage source in parallel, this element does not affect
the value of the voltage source and the voltage across this element is determined by
the voltage source. In order to check whether the circuit is valid, we first simplify the
circuit by removing the elements in parallel with a voltage source, obtaining a
simplified circuit shown in Fig. P5.4.

Fig. S4.5
Appendix. Solutions to Exercise Problems 233

A valid circuit should obey the KVL, namely

ð3 VÞ þ ð1 VÞ þ ð3 VÞ ¼ 0,

which obviously is an incorrect equation. Thus, this circuit is invalid.

Problem 4.6 Use KVL to verify if the following circuit is valid.

Fig. P4.6

Solution
This circuit contains a controlled voltage source labeled by “2i,” where “i” is the
voltage drop across the 1 Ω resistor. This controlled source is referred to as a current-
controlled voltage source. The unit of “i” is amperes (A) and the unit of “2i” is volts
(V). Clearly, the conversion factor “2” has a unit of “volts per ampere” (V/A).

Voltage across the 1 Ω resistor is 3 V provided by the 3 V voltage source. Using


Ohm’s law, i can be obtained by the product (3 V)/(1 Ω) ¼ (3 A).  
The current-controlled voltage source then has a value of 2i ¼ 2 V A  ð3 A Þ ¼
6 V.
The KVL equation can be readily obtained as
234 Appendix. Solutions to Exercise Problems

ð3 VÞ þ ð3 VÞ ¼ 6 V,

which is valid. Therefore, this is a valid circuit. When we set up this KVL equation,
we only pay attention to the voltage sources because the elements connected to the
voltage sources in parallel do not affect the values of the voltage sources. We can
ignore other elements in parallel with the voltage sources when setting up equations
using Fig. S4.6, where the controlled voltage source is replaced by a regular voltage
source of 6 V after we figure out the value of the controlled source to be 6 V.

Fig. S4.6

Chapter 5. Kirchhoff’s Current Law (KCL)

Problem 5.1 Use KCL to verify if the following circuit is valid.

Fig. P5.1
Appendix. Solutions to Exercise Problems 235

Solution
Let us simplify the circuit by removing all voltage sources connected in series with a
current source because if an element is connected in series with a current source, this
element can be ignored. Figure P5.1 is then simplified as Fig. S5.1.

Fig. S5.1

There are two nodes in the circuit, and we can set up a KCL equation at one of
these nodes. The total current running into the node should equal to the total current
running out from the node. This leads to

ð1 AÞ þ ð3 AÞ ¼ 2 A,

which is clearly not valid. Therefore, the circuit is invalid.

Problem 5.2 Use KCL to verify if the following circuit is valid.

Fig. P5.2
236 Appendix. Solutions to Exercise Problems

Solution
This circuit contains a controlled current source labeled by “4v,” where “v” is the
voltage drop across the 1 Ω resistor. This controlled source is referred to as a voltage-
controlled current source. The unit of “v” is volts (V) and the unit of “4v” is amperes
(A). Clearly, the conversion factor “4” has a unit of “amperes per volt” (A/V).

The current through the 1 Ω resistor is 3 A provided by the 3 A current source.


Using Ohm’s law, v can be obtained by the product (1 Ω) (3 A) ¼ 3 V.
The voltage-controlled current source then has a value of 4v ¼ 4
(A/V)  (3 V) ¼ 12 A.
The KCL equation can be readily obtained as

ð9 AÞ þ ð3 AÞ ¼ 12 A,

which is valid. Therefore, this is a valid circuit. When we set up this KCL equation,
we only pay attention to the current sources because the elements connected to the
current sources in series do not affect the values of the current sources. If a circuit
branch contains a current source, the current in that branch is determined only by the
current source, and we can ignore other elements in that branch when setting up
equations.

Problem 5.3 Set up the KCL equations for the following Wheatstone bridge circuit.

Fig. P5.3

Solution
This circuit has four nodes, and we can set up three KCL equations at nodes a, b,
and c.
Appendix. Solutions to Exercise Problems 237

a : i1 þ i2 ¼ 5 A,

b : i1 ¼ i3 þ i5 ,

c : i2 þ i5 ¼ i4 :

We do not need to set up a KCL equation at node d, because this equation does
not provide any new information about the circuit. In fact, if we substitute the
equation at node b and the equation at node c into the equation at node a, we get

i3 þ i4 ¼ 5 A,

which is the KCL equation at node d.


In general, if the circuit has n nodes, we can only get n-1 independent KCL
equations.

Problem 5.4 Find the current i1 in the circuit shown in Fig. P5.4.

Fig. P5.4

Solution
This circuit has three nodes, and we would normally set up two KCL equations.
However, if a node is directly connected to a voltage source, we do not set up a KCL
equation for that node. Likewise, if a loop contains a current source, we do not set up
a KVL equation for that loop.

We only need to set up one KCL equation at node a as

ð2iÞ þ i1 ¼ i þ ð1 AÞ,

that is,
238 Appendix. Solutions to Exercise Problems

i þ i1 ¼ 1 A:

In this equation, we have two unknowns i and i1. We do not have enough
equations to solve for the unknown i1.
There are three meshes in the circuit, and we normally set up two KVL equations
for this circuit. However, two of the three meshes contains a current source. One
mesh contains an independent 1 A current source and the other mesh contains a
controlled (dependent) “2i” current source. We can only set up one KVL in the loop
that contains a 1 V voltage source and two resistors.
Using Ohm’s law, the voltage drop across the 1 Ω resistor is (1 Ω) (i1), and the
voltage drop across the 2 Ω resistor is (2 Ω) (i). Thus, the KVL equation is

ð1 ΩÞ  i1 þ ð2 ΩÞ  i ¼ 1 V:

We now have a system of two equations with two unknowns i and i1.

i þ i1 ¼ 1 A
ð1 ΩÞ  i1 þ ð2 ΩÞ  i ¼ 1 V:

Divided by (1 Ω) on both sides of the second equation, the system becomes



i þ i1 ¼ 1 A
2i þ i1 ¼ 1 A:

This system has a solution of



i¼0A
i1 ¼ 1 A:

Problem 5.5 This circuit model a transistor, which has many applications such as
amplifiers. Find ib.
Appendix. Solutions to Exercise Problems 239

Fig. P5.5

Solution
We need to use both KVL and KCL for this problem.

There are three meshes in the circuit, and we normal can set up three KVL
equations. The upper-left mesh contains a controlled current source, and we do not
set up a KVL equation for that mesh.
There are four nodes in the circuit, and we set up three KCL. We generally want
to avoid KCL equations that involve voltage sources. However, ib is what we are
looking for and ib is used in the controlled source calculation. This problem forces us
to set up equations that include ib. Therefore, the KCL equations at nodes b and c are
set up. We thus have a system of four equations:
8
>
> KVL : R1 i1 þ R2 i2 ¼ V cc
>
>
< KVL : R i ¼ V þ R i
2 2 b e 3
>
> KCL at b : i1 ¼ i2 þ ib
>
>
:
KCL at c : ib þ βib ¼ i3

Next, we substitute the KCL equations into KVL equations and obtain
240 Appendix. Solutions to Exercise Problems


KVL : R1 ði2 þ ib Þ þ R2 i2 ¼ V cc
KVL : R2 i2 ¼ V b þ Re ð1 þ βÞib

Rearranging the terms yields



ðR1 þ R2 Þi2 ¼ V cc  R1 ib
R2 i2 ¼ V b þ Re ð1 þ βÞib

Eliminating i2 yields

V cc  R1 ib V b þ Re ð1 þ βÞib
¼
R1 þ R2 R2
Finally,

V cc
 VR2b
ib ¼ R ðR1þβ
1 þR2
Þ R1
:
e
R2 þ R1 þR 2

Chapter 6. Resistors in Series and in Parallel

Problem 6.1 Ten 1 kΩ resistors are connected in series, the total resistance is

(a) 10 Ω
(b) 100 Ω
(c) 1 kΩ
(d) 10 kΩ
(e) 100 kΩ

Solution
(d) 10 kΩ
For ten 1 kΩ resistors in series, the total resistance is

Rtotal ¼ 10  ð1 kΩÞ ¼ 10 kΩ:

Problem 6.2 Ten 1 kΩ resistors are connected in parallel, the total resistance is

(a) 10 Ω
(b) 100 Ω
(c) 1 kΩ
(d) 10 kΩ
(e) 100 kΩ
Appendix. Solutions to Exercise Problems 241

Solution
(b) 100 Ω
For ten 1 kΩ resistors in parallel, the total resistance is

1 1 1
¼ 10  ¼ :
Rtotal 1 kΩ 100 Ω
Thus,

Rtotal ¼ 100 Ω:

Problem 6.3 Two resistors R1 and R2 are connected in series. The total resistance is
1 kΩ.

(a) R1 is less than 1 kΩ.


(b) R1 is larger than 1 kΩ.
(c) R1 is 500 Ω.
(d) R1 and R2 must have the same resistance.
(e) R1 and R2 must not have the same resistance.

Solution
(a) R1 is less than 1 kΩ.
When R1 and R2 are connected in series, the total resistance is

Rtotal ¼ R1 þ R2 ¼ 1 kΩ:

Since R1 and R2 are positive, none of them can be larger than 1 kΩ.

Problem 6.4 Two resistors R1 and R2 are connected in parallel. The total resistance
is 1 kΩ.

(a) R1 is less than 1 kΩ.


(b) R1 is larger than 1 kΩ.
(c) R1 is 2 kΩ.
(d) R1 and R2 must have the same resistance.
(e) R1 and R2 must not have the same resistance.

Solution
(b) R1 is larger than 1 kΩ.
When resistors are connected in parallel, the total resistance is smaller than the
smallest resistor. If the total resistance is 1 kΩ, then both R1 and R2 must be larger
than 1 kΩ.

Problem 6.5 Four resistors R1, R2, R3, and R4 are connected in parallel. They satisfy
the relationship: R1 ¼ R2 < R3 ¼ R4. The total resistance is 1 kΩ.
242 Appendix. Solutions to Exercise Problems

(a) R1 is less than 1 kΩ.


(b) R1 is less than 2 kΩ.
(c) R3 is less than 1 kΩ.
(d) R3 is less than 2 kΩ.
(e) R3 is less than 4 kΩ.
(f) R1 is larger than 4 kΩ.
(g) None of the above.

Solution
(e) None of the above.

The total resistance is calculated as

1 1 1 1 1 2 2
¼ þ þ þ ¼ þ :
Rtotal R1 R2 R3 R4 R1 R3
Let us consider the constraints for R1.
It is given that Rtotal ¼ 1 kΩ and R1 < R3, and then

1 2 2 2 2 4
¼ þ < þ ¼ :
1kΩ R1 R3 R1 R1 R1
That is,

1 1
< :
4 kΩ R1
R1 < 4 kΩ:

On the other hand, R3 > 0, then

1 2 2 2
¼ þ > :
1 kΩ R1 R3 R1
That is,

1 1
> :
2 kΩ R1
R1 > 2 kΩ:

Finally,

2 kΩ < R1 < 4 kΩ:

Let us consider the constraints for R3.


Since R1 < R3,
Appendix. Solutions to Exercise Problems 243

1 2 2 2 2 4
¼ þ > þ ¼ :
1kΩ R1 R3 R3 R3 R3
That is,

1 1
> :
4 kΩ R3
R3 > 4 kΩ:

Problem 6.6 Ten resistors are connected in parallel, with Rn ¼ n Ω, for n ¼ 1, 2, . . .,


10.

(a) Rtotal ¼ 1 + 2 + 3 + 4 + 5 + 6 + 7 + 8 + 9 + 10 ¼ 55 Ω


(b) Rtotal ¼ 11 þ 12 þ 13 þ 14 þ 15 þ 16 þ 17 þ 18 þ 19 þ 10
1
¼ 2:9290 Ω
(c) Rtotal < 1 Ω
(d) Rtotal ¼ 1þ2þ3þ4þ5þ6þ7þ8þ9þ10
12345678910
Ω

Solution
(c) Rtotal < 1 Ω

The total resistance is calculated as

1 1 1 1 1 1 1 1 1
¼ þ þ þ ... þ ¼ þ þ þ ... þ :
Rtotal R1 R2 R3 R10 1 2 3 10
1 1
Rtotal ¼ ¼ :
1
R1 þ R12 þ R13 þ . . . þ R110 11 þ 12 þ 13 þ . . . þ 10
1

1
Rtotal < ¼ 1 Ω:
1
10 þ 1
10 þ 1
10 þ . . . þ 10
1

The total resistance is always smaller than the smallest resistor when the resistors
are connected in parallel.

Problem 6.7 Find the total resistance for the resistor network shown in Fig. P6.7.
Each resistor in the network is 1 Ω.

Fig. P6.7
244 Appendix. Solutions to Exercise Problems

Solution
This resistor network can be divided into three sections. The right section contains
four resistors connected in parallel as shown in Fig. S6.7a. Since all four resistors are
the same, and each has a resistance of 1 Ω. The equivalent resistance is 1/4 Ω.

Fig. S6.7a

Similarly, the left section contains two resistors connected in parallel as shown in
Fig. S6.7b. These two resistors have the same value of 1 Ω. Thus, the equivalent
resistance of these two resistors is 1/2 Ω.

Fig. S6.7b

Finally, these three sections can be represented as three resistors connected in


series as shown in Fig. S6.7c. The total resistance is the sum

Rtotal ¼ ð0:5 ΩÞ þ ð1 ΩÞ þ ð0:25 ΩÞ ¼ 1:75 Ω:

Fig. S6.7c

Chapter 7. Voltage Divider and Current Divider

Problem 7.1 In the circuit shown in Fig. P7.1, R1 ¼ 1 kΩ. Find the values of other
resistors.
Appendix. Solutions to Exercise Problems 245

Fig. P7.1

Solution
This is a voltage divider circuit.

Resistors R1 and R2 divide the voltage in half. In other words, the voltage drop on
R1 is the same as the voltage drop on R2. This implies that R2 ¼ R1 ¼ 1 kΩ.
The voltage drop on R3 is 40 V, which is four times the voltage drop on R1.
Therefore,

R4 ¼ 4R1 ¼ 4  1 kΩ ¼ 4 kΩ:

We must point out that the voltage divider technique can only be applied when
the resistors involved have the same current. Do not apply the voltage divider
technique if the resistors have different currents. In our problem, the voltage
terminals labeled “10 V,” “20 V,” and “60 V” do not connect to other elements. If
other elements are connected to the voltage terminals, the voltage divider technique
cannot be used.
246 Appendix. Solutions to Exercise Problems

Problem 7.2 Find the voltage v in the circuit shown in Fig. P7.2. All resistors have
the value of 1 Ω.

Fig. P7.2

Solution
First of all, R1 is in series of the current source, and it does not affect the current of
the source. Therefore, R1 can be removed without affecting the rest of the circuit.

Resistors R4, R5, and R6 are in parallel, and each of them has a value of 1 Ω. The
combined resistor is labeled as R456. The value of these three resistors is 1/3 Ω.
Resistors R2 and R3 are in series, and each of them has a value of 1 Ω. The
combined value of these two resistors is labeled as R23. The value of R23 is 2 Ω.
Figure P7.2 can be reduced to Fig. S7.2a.

Fig. S7.2a
Appendix. Solutions to Exercise Problems 247

Figure S7.2 is further reduced to Fig. S7.2b, by combining resistors R23 and R456.
We denote the combed resistor as R23456. Since R23 ¼ 2 Ω and R456 ¼ 1/3 Ω,
R23456 ¼ R23 + R456 ¼ 2 + 1/3 ¼ 7/3 Ω.

Fig. S7.2b

The circuit in Fig. S7.2b is a typical current divider, the current i1 can be
calculated as

R7 ð 1 ΩÞ 3
i 1 ¼ ð1 A Þ ¼ ð1 AÞ ¼ A:
R7 þ R23456 ð1ΩÞ þ ð7=3 ΩÞ 10

Finally, using Ohm’s law, we have (see Fig. S7.2a)


   
3 1 1
v ¼ i1  R456 ¼ A  Ω ¼ V ¼ 0:1 V:
10 3 10

Problem 7.3 Calculate currents i1, i2, i3, and i4 in the circuit shown in Fig. P7.3.

(a) R1 ¼ R2 ¼ R3 ¼ R4 ¼ 1 Ω.
(b) R1 ¼ 1 Ω, R2 ¼ 2 Ω, R3 ¼ 3 Ω, and R4 ¼ 4 Ω.

Fig. P7.3
248 Appendix. Solutions to Exercise Problems

Solution
(a) The condition R1 ¼ R2 ¼ R3 ¼ R4 ¼ 1 Ω leads to

10 A
i1 ¼ i2 ¼ i3 ¼ i4 ¼ ¼ 2:5 A:
4
If all the resistors are of the same value, they share the same amount of current.
(b) We need to use the current divider formula four times for this part.

1 1
R1 R1 24 1 24
i1 ¼ ð10 AÞ ¼ ð10 AÞ 1 ¼  ¼ A
1
R1 þ R12 þ R13 þ R14 1þ2þ3þ4
1 1 1 5 R1 5

1 1
R2 R2 24 1 12
i2 ¼ ð10 AÞ ¼ ð10 AÞ 1 ¼  ¼ A
1
R1 þ 1
R2 þ 1
R3 þ 1
R4 1 þ þ þ
1
2
1
3
1
4
5 R2 5

1 1
R3 R3 24 1 8
i3 ¼ ð10 AÞ ¼ ð10 AÞ 1 ¼  ¼ A
1
R1 þ R12 þ R13 þ R14 1þ2þ3þ4
1 1 1 5 R3 5

1 1
R4 R4 24 1 6
i4 ¼ ð10 AÞ ¼ ð10 AÞ 1 ¼  ¼ A
1
R1 þ R12 þ R13 þ R14 1þ2þ3þ4
1 1 1 5 R4 5

Problem 7.4 Find i1 and i2 in the circuit shown in Fig. P7.4.

Fig. P7.4

Solution
The left section of the circuit is a voltage divider. The resistor gets its voltage
proportional to it resistance value. The ratio of R2:R1 is 3:2. The voltage source
Appendix. Solutions to Exercise Problems 249

distributes 5 V to R2 and R1 according to this ratio. As a result, R2 gets 3 volts, and R1


gets 2 volts. Therefore,

v ¼ 3 V:

The right section of the circuit is a current divider. The voltage-controlled current
source generates a current of 2v ¼ 6 A.

For a current divider, the total current is distributed to the resistors according to
the ratio of conductance. The conductance is the reciprocal of the resistance. The
total current of 6A is distributed to R3 and R4 according to the ratio

1 1 1 1
: ¼ : ¼ 1 : 5:
R3 R4 5 1
In other words, R4 gets five times more current than R3 gets. Therefore,

i1 ¼ 1 A,

i2 ¼ 5 A:

For a voltage divider, a larger resistor gets a larger portion of the voltage.
For a current divider, a larger resistor gets a smaller portion of the current.
You may notice that there is a wire connecting the left section and the right
section in this circuit. There is no current in this wire because current can only flow
in a complete loop, and this wire is not part of a loop.

Problem 7.5 For a voltage divider circuit shown in Fig. P7.5. R1 is 1 Ω. The voltage
across R1 is 1 V. Is it possible to determine the source voltage v and the value of the
other resistor R2?

Fig. P7.5
250 Appendix. Solutions to Exercise Problems

Solution
Using the voltage divider formula, we have

R1 ð1ΩÞ
1V¼v ¼v :
R1 þ R2 ð1ΩÞ þ R2

We only one equation but two unknowns v and R2. We are unable to obtain a
unique solution. In fact, we can have infinite number of solutions. Here are three
solutions:

R2 ¼ 1 Ω, v ¼ 2 V;

R2 ¼ 2 Ω, v ¼ 3 V;

R2 ¼ 3 Ω, v ¼ 4 V:

Chapter 8. Node-Voltage Method

Problem 8.1 Set up node equations for the circuit given in Fig. P8.1.

Solution
The ground is selected as the reference node. We will calculate the voltages at the
other two nodes, noted as v1 and v2. The two node equations are set up as follows.

Fig. P8.1
Appendix. Solutions to Exercise Problems 251

v1  v v1 v1  v2
þ þ ¼ 0,
R1 R3 R2
v2  v1 v2
þ ¼ i:
R2 R4
Here, R1, R2, R3, R4, v, and i are given. The two unknowns are v1 and v2. The node
equations are essentially KCL equations. Each term in the node equations is a
current.

Problem 8.2 Set up node equations for a circuit containing a controlled source.

Fig. P8.2

Solution
The procedure is almost the same as in Problem 8.1, except that one extra equation is
needed for the relationship regarding the controlled source.

The ground is selected as the reference node. We will calculate the voltages at the
other two nodes, noted as v1 and v2. The two node equations are set up as follows.
v1  v v1 v 1  v2
þ þ ¼ 0:
R1 R3 R2
252 Appendix. Solutions to Exercise Problems

v2  v1 v2
þ ¼ 2i:
R2 R4
We need one more equation for the controlled source relationship:

v1  v2
i¼ :
R2
Here, R1, R2, R3, R4, and v are given. The unknowns are v1 and v2.

Problem 8.3 Set up the node equations for a circuit, in which a voltage source is
between the two nodes, and there are no resistors between these two nodes.

Fig. P8.3

Solution
If there is nothing but a voltage source between two nodes, the regular node
equations cannot be established because we do not know how to express the current
between these two nodes. A well accepted remedy is to use a super node
(as indicated by a dotted ellipse in Fig. P8.3).

We will only have one unknown, v2. The node voltage v1 will be expressed as
v2 + 10 V. The following is the node equation for the super node:
Appendix. Solutions to Exercise Problems 253

v2 þ 10  v v2 þ 10 v2
þ þ ¼ i:
R1 R3 R4

Problem 8.4 Set up the node equations for a circuit, in which a controlled voltage
source is between the two nodes, and there are no resistors between these two nodes.

Fig. P8.4

Solution
If there is nothing but a controlled voltage source between two nodes, we will use a
super node (as indicated by a dotted ellipse in Fig. P8.4). The controlled source
needs an extra equation to describe the controlling information. We need to define
“i” to calculate “2i.”

We will only have one unknown, v1. The node voltage v2 will be expressed as
v2 + 2i. The following is the node equation for the super node:
v1  v v1 v1 þ 2i
þ þ ¼ i0 :
R1 R3 R4
We need one more equation for the controlled source relationship:
254 Appendix. Solutions to Exercise Problems

v1
i¼ :
R3
There is not much difference in setting up equation for independent sources and
for controlled sources. When a controlled source is used, an extra equation is
required to describe the depending variable.

Problem 8.5 Set up node equations for the circuit, where a voltage source is
between a node and the reference node.

Fig. P8.5

Solution
In this circuit, a voltage source is between a node and the reference node. In this case,
the voltage of v1 is known, and there is no need to set up a node equation for v1. We
only need one node equation for v2, which is the only unknown to be solved.

v2  v v2
þ ¼ i:
R2 R4

Chapter 9. Mesh-Current Method

Problem 9.1 Set up the mesh equations and solve for the mesh currents.
Appendix. Solutions to Exercise Problems 255

Fig. P9.1

Solution
Two meshes have two mesh equations. Let the mesh current for the left mesh be i1
and the mesh current for the right mesh be i2. We have

ð1 ΩÞði1 Þ þ ð2 ΩÞði1  i2 Þ ¼ 1 V,

ð3 ΩÞði2 Þ þ ð2 ΩÞði2  i1 Þ ¼ 2 V:

Rearranging the terms yields

3i1  2i2 ¼ 1,

2i1 þ 5i2 ¼ 2:

Let us solve this system of equations by hand. Multiplying the first equation by
2 and multiplying the second equation by 3, the above two equations become

6i1  4i2 ¼ 2,

6i1 þ 15i2 ¼ 6:

Summing these two equations leads to

11i2 ¼ 4:
4
i2 ¼  A:
11
Current i1 can be solved from 3i1  2i2 ¼ 1,
 
4
3i1  2  ¼ 1,
11
256 Appendix. Solutions to Exercise Problems

8 3
3i1 ¼ 1  ¼ ,
11 11
13
i1 ¼ A:
11

Problem 9.2 This circuit contains a voltage-controlled voltage source. Set up the
mesh equations and solve for the mesh currents.

Fig. P9.2

Solution
We can treat the controlled source as a regular source and set up the mesh equations.

Two meshes have two mesh equations. Let the mesh current for the left mesh be i1
and the mesh current for the right mesh be i2. We have

ð1 ΩÞði1 Þ þ ð2 ΩÞði1  i2 Þ ¼ 1 V,

ð3 ΩÞði2 Þ þ ð2 ΩÞði2  i1 Þ ¼ 2v:

The controlled source needs the value v, which can be calculated by

v ¼ ð1 ΩÞði1 Þ,

2v ¼ 2i1 :

Then the two mesh equations can be rewritten as

ð1 ΩÞði1 Þ þ ð2 ΩÞði1  i2 Þ ¼ 1 V,

ð3 ΩÞði2 Þ þ ð2 ΩÞði2  i1 Þ ¼ 2i1 :

Rearranging the terms yields


Appendix. Solutions to Exercise Problems 257

3i1  2i2 ¼ 1,

4i1 þ 5i2 ¼ 0:

Let us solve this system of equations by hand. Multiplying the first equation by
4 and multiplying the second equation by 3, the above two equations become

12i1  8i2 ¼ 4,

12i1 þ 15i2 ¼ 0:

Summing these two equations leads to

7i2 ¼ 4:
4
i2 ¼ A:
7
Current i1 can be solved from 4i1 + 5i2 ¼ 0.
 
4
4i1 þ 5 ¼ 0,
7
20
4i1 ¼ ,
7
5
i1 ¼ A:
7

Problem 9.3 This circuit contains a current source. Set up the mesh equations and
solve for the mesh currents.

Fig. P9.3
258 Appendix. Solutions to Exercise Problems

Solution
Since we do not know how to express the voltage drop across a current source, we
want to avoid current source in our mesh equations. Using a super mesh can avoid
the current sources. For this problem, a super mesh is indicated in Fig. S9.3 as a
dotted loop.

Fig. S9.3

In Fig. S9.3, i1 and i2 are related as

i2  i1 ¼ 1 A:

We have only one mesh equation for this super mesh:

ð1ΩÞði1 Þ þ ð3ΩÞði2 Þ ¼ 1 V  2 V:

Combining the two above equations yields

i1 þ 3ð1 þ i1 Þ ¼ 1,

4i1 ¼ 4,

i1 ¼ 1 A:

Using i2  i1 ¼ 1 A, we have

i2 ¼ 0 A:
Appendix. Solutions to Exercise Problems 259

Problem 9.4 What if the current source is a controlled current source?

Fig. P9.4

Solution
We use exactly the same strategy to deal with a controlled current source as with an
independent current source. Using a super mesh is able to avoid a current source. The
dotted loop in Fig. P9.4 indicates a super mesh, which does not contain the
controlled current source. The controlling variable is v, which can be calculated
via Ohm’s law as

v ¼ ð3 ΩÞði2 Þ:

In Fig. P9.4, i1 and i2 are related as

i2  i1 ¼ 3v ¼ 9i2 ,

i1 ¼ 8i2 :

We have only one mesh equation for this super mesh:

ð1 ΩÞði1 Þ þ ð3 ΩÞði2 Þ ¼ 1 V  2 V:

Combining the two above equations yields

8i2 þ 3i2 ¼ 1,

5i2 ¼ 1,
1
i2 ¼ A:
5
Using i1 ¼  8i2, we have
260 Appendix. Solutions to Exercise Problems

8
i1 ¼  A:
5

Problem 9.5 Application of the mesh equations, considering a special case of a


current course in the circuit.

Fig. P9.5

Solution
This current source is at a special location. The mesh current is the same as the
current source value. In this case, the mesh current is already determined. There is no
need to set up an equation for the mesh current i1. We already know that

i1 ¼ 1 A:

We only need one mesh current equation in terms of i2.

ð3 ΩÞði2 Þ þ ð2 ΩÞði2  1Þ ¼ 2 V:

Rearranging the terms yields

2 þ 5i2 ¼ 2,

5i2 ¼ 4,
4
i2 ¼  A:
5

Chapter 10. Computer Simulation Software Multisim

Problem 10.1 Use Multisim to simulate a circuit shown to determine the node
voltage. You can choose any resistor values.
Appendix. Solutions to Exercise Problems 261

Fig. P10.1

Solution
Run Multisim on a computer. We will create a new project.

File ! New ! Blank


Place ! Component
A new window pops out for you to select components.
To find resistors, you can type “resistor” in the search area or select “Basic” under
“Group.”

After clicking “OK,” a resistor will be placed in the workplace.


To find voltage sources, you can type “DC_POWER” in the search area or select
“Sources” under “Group.”
After clicking “OK,” a dc voltage source will be placed in the workplace.
262 Appendix. Solutions to Exercise Problems

You need three resistors and two voltage sources.

To rotate a component, you right-click and select “Rotate 90 .”


To wire up the circuit, go to Place ! Wire and connect the components one
by one.

To change the value of a voltage source, double-click or right-click the compo-


nent and then select “Properties”. Change the voltage and click “OK.”
Appendix. Solutions to Exercise Problems 263

Select a multimeter from the right column. Connect to multimeter to the circuit.
264 Appendix. Solutions to Exercise Problems

Find a “GROUND” by going to Place ! Components. Under “Group,” select


“Sources.” Put two GROUNDs in the circuit as shown. The simulation will not run if
the circuit does not have a ground.

Click the multimeter and configure it to be a voltmeter.


To simulate the circuit and find the node voltage, go to Simulate ! Run
(or simply click the green triangle on the tool bar). You will see the result of the
simulation in the voltmeter.

To stop the simulation, go to Simulate ! Stop (or simply click the red square on
the tool bar).
Appendix. Solutions to Exercise Problems 265

Now, change some resistor values and repeat the experiment.

Problem 10.2 Chapter 15 discusses operational amplifiers (Op-Amps). Simulate an


inverting amplifier with a sinewave input. Change the power supply values to
observe any potential clipping in the output. The op-amp circuit is given in
Fig. P10.2.

Fig. P10.2

Solution
Run Multisim on a computer. We will create a new project.

File ! New ! Blank


266 Appendix. Solutions to Exercise Problems

Place ! Component
A new window pops out for you to select components.
To find resistors, you can type “resistor” in the search area or select “Basic” under
“Group.”

After clicking “OK,” a resistor will be placed in the workplace. You need two
resistors.
To find voltage sources, you can type “DC_POWER” in the search area or select
“Sources” under “Group.”
After clicking “OK,” a dc voltage source will be placed in the workplace. You
need two DC voltage sources. You need to flip one dc voltage source upside down
by right-clicking it and rotating it 90 twice.
To find an AC (alternating current) source, you can type “AC_POWER” in the
search area or select “Sources” under “Group.”
After clicking “OK,” an AC voltage source will be placed in the workplace.
An op-amp can be found by searching in Group ! Analog ! OPAMP.
There are many options to choose from. You can pick, for example, “741.” Click
“OK.”
Get two GROUNDs under “Sources.”
Change the values of the components by clicking the component and inputting
the correct values, which are shown in the following schematic.
The op-amp wiring is rather confusing. The co-amp is first flipped upside down.
The inverting input () “pin 2” is connected to the two resistors. The non-inverting
input (+) “pin 3” is connected to the ground. The output “pin 6” is connected to the
feedback resistor. A positive power supply is connected to “pin 7.” A negative power
supply is connected to “pin 4.”
Appendix. Solutions to Exercise Problems 267

An oscilloscope with two inputs can be found on the right column. “Input A” of
the oscilloscope is used to monitor the op-amp output. “Input B” of the oscilloscope
is used to monitor the op-amp input.

Clicking on the oscilloscope symbol opens the oscilloscope display.


Run a simulation by clicking the green triangle on the tool bar. In order to better
visualize the waveforms in the oscilloscope display, you are encouraged to adjust the
scaling setting for Timebase, Channel A, and Channel B as indicated in the follow-
ing figure.
In the oscilloscope display, the red wave is the output and the green wave is the
input. The default color is red. To get a green curve, you need to right-click the wire
that is attached to the input source. Selecting the “Property” option. Change the “Net
color” to green.
268 Appendix. Solutions to Exercise Problems

Next, we change the negative power supply to 0 V as indicated below. An


equivalent way is to connect the op-amp’s “pin4” to the ground. After running
another simulation with this change, we observed that the output is clipped.
Appendix. Solutions to Exercise Problems 269

Finally, we change both voltage sources to 1 V and run a simulation. The results
are shown below. The output waveform is clipped on both positive and negative
directions.
270 Appendix. Solutions to Exercise Problems

Problem 10.3 Operational amplifier circuits are normally designed to operate from
dual supplies, e.g., +9 V and  9 V. This is not always easy to achieve and therefore
it is often convenient to use a single-ended or single supply version of the electronic
circuit design. Find a single supply op-amp circuit and use Multisim circuit.

Solution
The following circuit uses only one 9 V power supply. It can amplify a sinewave
signal by introducing a positive bias. The output is an amplified signal with a
positive dc bias. The Multisim circuit and simulation result are shown below.
Appendix. Solutions to Exercise Problems 271
272 Appendix. Solutions to Exercise Problems

Chapter 11. Superposition

Problem 11.1 A student uses the superposition principle to solve the voltage v. The
student’s answer is wrong. Please help this student to find the mistake. Let us start
with Fig. P11.1a.

Fig. P11.1a

There are two sources in the circuit.


Case 1: Let us first remove the source on the left, obtaining Fig. P11.1b. It can be
verified that the solution is v ¼ 0.

Fig. P11.1b

Case 2: Let us first remove the source on the right, obtaining Fig. P11.1c. The left
two resistor are in parallel. Therefore, these two right resistors can be combined into
a 0.5 k resistor, as shown in Fig. P11.1c.
Appendix. Solutions to Exercise Problems 273

Fig. P11.1c

Fig. P11.1d

This is a voltage divider, and we have v ¼ 2/3 V.


Combining the results from these two steps, the final answer is v ¼ 2/3 V.
However, the correct answer for v is 2 V.

Solution
The student’s approach is wrong. We cannot treat a controlled source as an indepen-
dent source. A controlled source can never be removed. This circuit has only one
independent source, and we cannot use the super position principle to solve this
problem.

Problem 11.2 Another student tries to use the superposition principle to solve for
the voltage v in a different problem. The student’s answer is wrong. Please help this
student to find the mistake. Let us start with Fig. P11.2a.
274 Appendix. Solutions to Exercise Problems

Fig. P11.2a

There are two independent sources in this circuit, and we will use the superposi-
tion principle to solve this problem.
Case 1: We remove the left source. The circuit becomes Fig. P11.2b.

Fig. P11.2b

The 1 k resistor on the right has no effect in the circuit. The two resistors on the
left is a current divider. Each 1 kΩ resistor on the left gets 0.5A of current. Using
Ohm’s law, v ¼ -500 V.
Case 2: We remove the right source. The circuit becomes Fig. P11.2c.

Fig. P11.2c
Appendix. Solutions to Exercise Problems 275

After combining the right two 1 kΩ resistors into a 0.5-kΩ resistor, we obtain a
voltage divider. In this case, v ¼ 2/3 V.
According to the superposition principle, we combine these two answers and
obtain the final answer of

2 1 1
v¼  ¼ V:
3 2 6
However, the correct answer is

v ¼ 499:5 V:

What is wrong?

Solution
In the second case, when we remove the current source, we should leave the circuit
open (instead of shorting the circuit).

Remember:
Removing a voltage source ¼ setting the voltage to 0 ¼ short.
Removing a current source ¼ setting the current to 0 ¼ open.

Problem 11.3 Use the superposition principle to solve for the voltage v. The circuit
is

Fig. P11.3

Solution
Case 1: Remove the left source by shorting the circuit and consider the circuit in
Fig. S11.3a.
276 Appendix. Solutions to Exercise Problems

Fig. S11.3a

We can set up a node equation for this circuit.

v v þ 3v
þ ¼ 1 A:
1 kΩ 1 kΩ
Thus,

v ¼ 1000 V:

Case 2: Remove the right source by leaving the circuit open and consider the
circuit in Fig. S11.3b.

Fig. S11.3b

There is only one loop in the circuit, and the current is the same everywhere in the
loop. The voltage rule applies. Each of the 1 kΩ resistor gets the same voltage. We
have a KVL equation

v þ v ¼ 3v þ ð1 VÞ:

Thus,
Appendix. Solutions to Exercise Problems 277

v ¼ 1 V:

Finally, combining results from Cases 1 and 2, we have

v ¼ 1000  1 ¼ 999 V:

Chapter 12. Thévenin and Norton Equivalent Circuit

Problem 12.1 Find the Thévenin equivalent circuit of the circuit in Fig. P12.1.

Fig. P12.1

Solution
To find RTh, we remove all the independent sources in the circuit, obtaining
Fig. S12.1.

The 2 Ω resistor and the 6 Ω resistor are in parallel, the combined resistor of them
is 1.5 Ω. Therefore,

RTh ¼ ð1 ΩÞ þ ð1:5 ΩÞ ¼ 2:5 Ω:

To find vTh, we just need to use a node equation to find the voltage v as indicated
in Fig. P12.1. Since there is no current in the 1 Ω resistor, vTh ¼ v. The node equation
is as follows.

v  ð1 VÞ v
þ ¼ 1 A,
2Ω 6Ω
278 Appendix. Solutions to Exercise Problems

Fig. S12.1

where the 1 Ω resistor is not considered because there is no current through


it. Solving this equation, we have

3v  ð3 VÞ v
þ ¼ 1 A,
6Ω 6Ω
4v ¼ 9V,
9
vTh ¼ v ¼ V:
4

Problem 12.2 Find the Thévenin equivalent circuit of the circuit in Fig. P12.2.

Fig. P12.2
Appendix. Solutions to Exercise Problems 279

Solution
To find vTh, we just need to use a node equation to find the voltage v as indicated in
Fig. P12.2. Since there is no current in the 1 Ω resistor, vTh ¼ v. The node equation is
as follows.

v  ð1 V Þ v
þ ¼ 3v,
2Ω 6Ω
where the 1 Ω resistor is not considered because there is no current through
it. Solving this equation, we have

3v  ð3 VÞ v
þ ¼ 3v,
6Ω 6Ω
3v  ð3 VÞ þ v ¼ 18v,

14v ¼ 3 V,
3
vTh ¼ v ¼  V:
14

Fig. S12.2

This circuit contains a controlled source; we cannot find RTh, by simply removing
the independent sources. We use a different approach. We short circuit the output
port and calculate the short-circuit current isc as labeled in Fig. S12.2.

We use the node equation to find the node voltage v and then use Ohm’s law to
find isc. The node equation for the circuit in Fig.S12.2 is given below.
280 Appendix. Solutions to Exercise Problems

v  ð1 V Þ v v
þ þ ¼ 3v,
2Ω 6Ω 1Ω
3v  ð3 VÞ v 6v
þ þ ¼ 3v,
6Ω 6Ω 6Ω
3v  ð3 VÞ þ v þ 6v ¼ 18v,

8v ¼ 3 V,
3
v¼ V:
8
Applying Ohm’s law on the 1 Ω resistor,

 38 V 3
isc ¼ ¼  A:
1Ω 8
Finally,

vTh  143 V 4
RTh ¼ ¼ ¼ Ω:
isc  38 A 7

Problem 12.3 Use the testing source method to find the Thévenin resistance RTh in
Problem 12.2.

Fig. P12.3
Appendix. Solutions to Exercise Problems 281

Solution
We first need to remove the independent source and add a test source at the output
port as shown in Fig. S12.3.

Fig. S12.3

Let us set up the node equation for the node voltage v.

v v v  vT
þ þ ¼ 3v,
2Ω 6Ω 1Ω
3v v 6v  6vT
þ þ ¼ 3v,
6Ω 6Ω 6Ω
3v þ v þ 6v  6vT ¼ 18v,

8v ¼ 6vT ,
3
v ¼  vT :
4
Applying Ohm’s law on the 1 Ω resistor,

vT  v 3 7
iT ¼ ¼ vT þ vT ¼ vT :
1Ω 4 4
Finally,
282 Appendix. Solutions to Exercise Problems

vT v 4
RTh ¼ ¼ 7 T ¼ Ω,
iT 4 v T 7

which is the same answer we obtain from Problem 12.2.

Problem 12.4 Find the Norton equivalent circuit of the circuit in Fig. P12.4.

Fig. P12.4

Solution
This circuit is identical to that in Problem 12.1. The Norton resistance is the same as
the Thévenin resistance. The methods for evaluating the Norton resistance are the
same for evaluating the Thévenin resistance. Therefore, we can directly use the result
from Problem 12.1 and the Norton resistance is

RNor ¼ RTh ¼ 2:5 Ω:

If we already know the Thévenin voltage, the Norton current can be readily
calculated as

vTh 9 V 9
iNor ¼ ¼ 54 ¼  A:
RNor 2 Ω
10

If the Thévenin voltage is not available, the Norton current can be calculated by
the short current at the output port (see Fig. S12.4).
Appendix. Solutions to Exercise Problems 283

Fig. S12.4

The node equation for v is

v  ð1 VÞ v v
þ þ ¼ 1 A,
2Ω 6Ω 1Ω
3v  ð3 VÞ v 6v
þ þ ¼ 1 A,
6Ω 6Ω 6Ω
10v ¼ 9 V,
9
v¼ V:
10
Using Ohm’s law on the 1 Ω resistor yields

9
v V 9
iNor ¼ ¼ 10 ¼ ¼ 0:9 A:
1Ω 1Ω 10
We reach the same answer.

Problem 12.5 Find the Norton equivalent circuit of the circuit in Fig. P12.5, using
the step-by-step Thévenin/Norton conversion method.
284 Appendix. Solutions to Exercise Problems

Fig. P12.5

Solution
In the first step, we convert the 1 V voltage source and the 2 Ω resistor into a Norton
equivalence, as shown in Fig. S12.5a, where the current source value is obtained by
(1 V)/(2 Ω) ¼ 0.5 A.

Fig. S12.5a

Combining the 2 Ω resistor and the 6 Ω resistor yields a 1.5-Ω resistor; combining
the 1 A source and the 0.5 A source yields a 1.5-A source (See Fig. S12.5b).
Appendix. Solutions to Exercise Problems 285

Fig. S12.5b

Next, we convert the 1.5 A current source and the 1.5 Ω resistor into a Thévenin
equivalence, as shown in Fig. S12.5c, where the voltage source value is obtained by
(1.5 A)(1.5 Ω) ¼ 2.25 V.

Fig. S12.5c

Combining the 1.5 Ω resistor and the 1 Ω resistor results in a 2.5-Ω resistor, as
shown in Fig. S12.5d, which is the Thévenin equivalent circuit of Fig. P13.1.

Fig. S12.5d
286 Appendix. Solutions to Exercise Problems

Finally, the Norton equivalent circuit is readily obtained from Fig. S12.5d as
Fig. S12.5e, where the Norton current value is calculated by (2.25 V)/
(2.5 Ω) ¼ 0.9 A.

Fig. S12.5e

Chapter 13. Maximum Power Transfer

Problem 13.1 Find the maximum power delivered to R in the circuit in Fig. P13.1
when R is set for maximum power transfer?

Fig. P13.1
Appendix. Solutions to Exercise Problems 287

Solution
From Problem 12.1, the Thévenin equivalent circuit has a Thévenin voltage of
2.25 V and a Thévenin resistance of 2.5 Ω (see Fig. S12.5d). Figure P13.1 is
equivalent to Fig. S13.1.

Fig. S13.1

According to the maximum power transfer principle, when R ¼ RTh ¼ 2.5 Ω, the
load R receives the maximum power.
The maximum power received by the load in this case is

ðvTh =2Þ2 ðð2:25 VÞ=2Þ2 81


pLoad ¼ ¼ ¼ ¼ 0:50625 W:
RTh 2:5 Ω 160

Problem 13.2 In Problem 13.1, let R ¼ 2.5 Ω. What is the power provided by the
1 V voltage source? What is the power provided by the 1 A current source? In the
circuit of Fig. P13.2, what percentage of the power delivered to the load R ¼ 2.5 Ω
by the two sources?
288 Appendix. Solutions to Exercise Problems

Fig. P13.2

Solution
To find the power of an individual power source, we have to use the original circuit,
instead of the Thévenin equivalent circuit.

The node equation for v is

v  ð1 VÞ v v
þ þ ¼ 1 A,
2Ω 6 Ω 1 Ω þ 2:5 Ω
21v  ð21 VÞ 7v 12v
þ þ ¼ 1 A,
42 Ω 42 Ω 42 Ω
21v  ð21 VÞ þ 7v þ 12v ¼ 42 V,

40v ¼ 63 V,
63
v¼ V:
40
Using Ohm’s law, the current running through the 2 Ω resistor from right to left is
63 
v  ð1 VÞ V  ð1 VÞ 23
i2Ω ¼ ¼ 40
¼ A:
2Ω 2Ω 80
This current is running into the positive terminal of the voltage source. This
means that this voltage is not providing any power to the circuit. Instead, this voltage
source is taking power from the circuit. In every life, this is the situation of a battery
is being charged. The power that is delivered from this voltage source is
 
23 23
pvoltage source ¼ ð1 VÞ A ¼ W:
80 80
Appendix. Solutions to Exercise Problems 289

For the 1A current source, the current is flowing out and pointing to v, which is
positive. This means that this source is providing power to the circuit. The voltage
across the current source is v ¼ 63/40 V. The power the is delivered from this voltage
source is
 
63 63
pcurrent source ¼ ð1 AÞ V ¼ W:
40 40

The current source gives 23/80 W to charge the voltage source and gives
   
63 23 103
W  W ¼ W
40 80 80

to the rest of the circuit. This 103/80 W is the total delivered power by the sources.
According to the result of Problem 13.1, the load consumes 81/160 W. The
percentage of the power delivered to the load R ¼ 2.5 Ω by the two sources is

81
81
160
103
¼ ¼ 39:3%:
80
206

We notice that under the maximum power transfer condition that the load
resistance equals to the Thevenin resistance, the percentage of the power delivered
to the load may not be 50%.

Problem 13.3 As shown by the result of Problem 13.2, when the load resistance
equals to the Thevenin resistance, the percentage of the power delivered to the load
can be less than 50%. Use an example to explain this phenomenon.

Solution
Let us consider a circuit in Fig. S13.3a and its Thévenin equivalent circuit
(Fig. S13.3b). The resistor R1 does not affect the Thévenin equivalent circuit.

Fig. S13.3a
290 Appendix. Solutions to Exercise Problems

Fig. S13.3b

When a load of R ¼ 1 Ω is attached to the output port a-b, the load will get a
maximum power of 1 W according to Fig. S13.3b.
We now calculate the power delivered from the 2 V voltage source. After a load
of R ¼ 1 Ω is attached to the output port a-b. Using Ohm’s law, the current running
through the two 1 Ω resistors is

2V
¼ 1 A:
ð 1 ΩÞ þ ð 1 ΩÞ

Also using Ohm’s law, the current running through the resistor R1 is

2V
:
R1
Therefore, the power delivered by the 2 V voltage source is

2V
pdeliverd ¼ ð1 AÞ þ  ð2 VÞ,
R1

which varies with the value of R1. The percentage of the power delivered to the
optimally matched load is given as

ploadd 1W
¼h i :
pdeliverd ð1 AÞ þ 2RV1  ð2 VÞ

When R1 ¼ 1,

ploadd 1W 1
¼ ¼ ¼ 50%:
pdeliverd ð1 AÞ  ð2 VÞ 2

When R1 ¼ 2 Ω,
Appendix. Solutions to Exercise Problems 291

ploadd 1W 1
¼ ¼ ¼ 25%:
pdeliverd ð1 AÞ þ 22 Ω
V
 ð2 VÞ 4

Chapter 14. Operational Amplifiers

Problem 14.1 Express the output voltage vout in terms of the inputs v1 and v2.

Fig. P14.1

Solution
This circuit has two inputs v1 and v2. We will use the superposition principle to find
the output vout.

Case 1: Let v2 ¼ 0. Figure P15.1 becomes Fig. S15.1.

Fig. S14.1a
292 Appendix. Solutions to Exercise Problems

There is no current at the non-inverting input. Thus, there is no current in the


resistors connecting to the non-inverting input. As a result, the voltage at the
non-inverting input is zero. Figure S14.1a is then reduced to Fig. S14.1b.

Fig. S14.1b

The circuit in Fig. S14.1b is an inverting amplifier. The output is evaluated as

100 kΩ
vout ¼  v ¼ 50v1 :
2 kΩ 1
Case 2: Let v1 ¼ 0. Figure P15.1 becomes Fig. S14.1c.

Fig. S14.1c
Appendix. Solutions to Exercise Problems 293

There is no current at the non-inverting input. The current in the two resistors
connected to the non-inverting is the same; therefore, we can use the voltage divider
method to find the voltage at the non-inverting input as

50 kΩ 1
vþ ¼ v ¼ v :
ð100 kΩÞ þ ð50 kΩÞ 2 3 2

Figure S15.3a is a non-inverting amplifier. Once v+ is known, the output can be


readily evaluated as
 
100 kΩ 51
vout ¼ 1þ vþ ¼ 51vþ ¼ v2 :
2 kΩ 3

According to the principle of superposition, we combine the results from Cases


1 and 2, obtaining

51
vout ¼ v  50v1 :
3 2

Problem 14.2 Consider a current source as the inverting input. Find the current
running into the output terminal.

Fig. P14.2

Solution
Since there is no current at the inverting input, the 10 kΩ resistor has the same
current as i1, running from right to left.

Using Ohm’s law, we have

v0 ¼ ð10 kΩÞði1 Þ:

The node equation for v0 is


294 Appendix. Solutions to Exercise Problems

v0
¼ i1 þ i0 :
2 kΩ
Combining the above two equations yields

ð10 kΩÞði1 Þ
¼ i1 þ i0 ,
2 kΩ
5i1 ¼ i1 þ i0 ,

i0 ¼ 4i1 :

Normally we are not interested in the current i0 at all in practice.


If you treat the op-amp as a super node, we may find it strange that there is no
current at the inverting input and the non-inverting input, but there is current at the
output. To solve this paradox, we need to realize that the op-amp has three other
terminals not commonly labeled in a circuit schematic: the positive power supply,
the negative power supply, and the ground. The current running into the output
terminal will go to the positive power supply, the negative power supply, or the
ground.

Problem 14.3 The circuit shown in Fig. P14.3 can be thought of a current source.
Find the range of the load RL, in which the current in RL is a constant. What is the
value of this constant current?

Fig. P14.3

Solution
Since there is no current at the non-inverting input terminal,

vþ ¼ 2 V:
Appendix. Solutions to Exercise Problems 295

Since there is no current at the inverting input terminal, the current in the 1 kΩ
resistor and the load resistor RL is the same. Thus, the voltage divider principle
applies, and we have

1 kΩ
v ¼ v:
RL þ ð1 kΩÞ 0

Since

vþ ¼ v ,

the current through the load RL is expressed as

v v 2V
i0 ¼ ¼ þ ¼ ¼ 2 mA:
1 kΩ 1 kΩ 1 kΩ
This constant current i0 ¼ 2 mA is maintained when the output voltage v0 in
between the rail voltages

0 < v0 < 10 V:

Since

RL þ ð1 kΩÞ R þ ð1 kΩÞ
0 < v0 ¼ v ¼ L ð2 VÞ < 10 V,
1 kΩ 1 kΩ
RL þ ð1 kΩÞ < 5 kΩ:

The condition for constant current is

0 < RL < 4 kΩ:

Chapter 15. Inductors

Problem 15.1 The switch closes at t ¼ 0. Find the inductor current iL as function
of time.

Fig. P15.1
296 Appendix. Solutions to Exercise Problems

Solution
We only know how to solve a circuit that contain only one inductor and one resistor.
Fig. P15.1 has two resistors!

To reduce this problem to a problem that we are able to solve, we treat the
inductor as the circuit load and find the Thévenin equivalent circuit of Fig. P16.1,
obtaining Fig. S15.1.

Fig. S15.1

At the initial time t ¼ 0,

iL ð0 Þ ¼ 0:

Since the inductor’s current cannot suddenly change,

iL ð0þ Þ ¼ 0:

At the final time t ¼ 1, the inductor can be treated as a conductor. Using Ohm’s
law,

1V
i L ð 1Þ ¼ ¼ 2 mA:
500 Ω
The time constant is calculated as

L 1H
τ¼ ¼ ¼ 0:002 s ¼ 2 ms:
R 500 Ω
Using the general mathematical expression of the inductor current

iL ðt Þ ¼ iL ð1Þ þ ½iL ð0Þ  iL ð1Þet=τ , for t  0,

we have
Appendix. Solutions to Exercise Problems 297

iL ðt Þ ¼ 2  2et=ð2msÞ mA, for t  0:

Problem 15.2 We use the same circuit as in Problem 15.1. We assume that the
switch has been closed for a long time. The switch opens at t ¼ 0. Find the inductor
current iL as function of time.

Fig. P15.2

Solution
After the switch opens, only one resistor on the right is effective in the circuit, and
the Thénenin equivalent circuit in Fig. S16.1 is no longer valid.

The initial condition of this problem is the final condition in Problem 15.1.
Using the result of Problem 15.1, at the initial time t ¼ 0,

iL ð0 Þ ¼ 2 mA:

Since the inductor’s current cannot suddenly change,

iL ð0þ Þ ¼ 2 mA:

At the final time t ¼ 1, the inductor can be treated a conductor and the current
diminishes to 0. That is,

iL ð1Þ ¼ 0:

The time constant is different from that in Problem 15.1 and is calculated as

L 1H
τ¼ ¼ ¼ 1 ms:
R 1 kΩ
Using the general mathematical expression of the inductor current
298 Appendix. Solutions to Exercise Problems

iL ðt Þ ¼ iL ð1Þ þ ½iL ð0Þ  iL ð1Þet=τ , for t  0,

we have

iL ðt Þ ¼ 2et=ð1msÞ mA, for t  0:

Problem 15.3 The switch in the circuit in Fig. P16.2 has been closed for a long time
before opening at t ¼ 0. Find the inductor’s current iL and the inductor’s voltage vL
for t  0.

Fig. P15.3

Solution
When the switch is closed for a long time, the inductor is a short circuit and the
inductor voltage vL ¼ 0. As a result, the controlled (dependent) source is also 0. In
other words, the dependent source is an open circuit. In this case, the inductor current
is same as the independent source 1 A. Thus,

iL ð0þ Þ ¼ iL ð0 Þ ¼ 1 A:

At time t ¼ 0, the switch opens, and Fig. P15.3 becomes Fig. S15.3a.

Fig. S15.3a
Appendix. Solutions to Exercise Problems 299

We do not know how to solve a circuit with an inductor and a dependent source.
Fortunately, our old friend “Thévenin equivalent circuit” is able to help.
To find the Thévenin equivalent circuit for the dependent source in Fig. S16.3a,
we apply a test source at the output port a-b, as indicated in Fig. S15.3b.

Fig. S15.3b

We have

iT ¼ 2vL ¼ 2vT,

and the Thévenin resistance can be calculated as

vT v
RTh ¼ ¼ T ¼ 0:5 Ω:
iT 2vT
To find the Thévenin voltage, we need to first find the short-circuit current as
shown in Fig. S15.3c.

Fig. S15.3c

The short circuit leads to vL ¼ 0, which implies 2vL ¼ 0 and isc ¼ 0. Thus, the
Thévenin voltage is

vTh ¼ isc  RTh ¼ 0:

Using the Thévenin equivalent circuit, Fig. S16.3b becomes Fig. S15.3d.
300 Appendix. Solutions to Exercise Problems

Fig. S15.3d

There is no source in Fig. S15.3d, and the inductor current will eventually
diminish to 0. Thus,

iL ð1Þ ¼ 0:

The time constant is calculated from Fig. S15.3d as

L 1H
τ¼ ¼ ¼ 2 s:
R 0:5 Ω
Using the general mathematical expression of the inductor current

iL ðt Þ ¼ iL ð1Þ þ ½iL ð0Þ  iL ð1Þet=τ , for t  0,

we have

iL ðt Þ ¼ et=ð1sÞ A, for t  0:

Finally, we find inductor voltage vL for t  0 according to the relationship

diL ðt Þ
vL ð t Þ ¼ L :
dt
Since L ¼ 1 H, and we have

det
vL ð t Þ ¼ ¼ et V, for t  0:
dt

Chapter 16. Capacitors

Problem 16.1 The switch closes at t ¼ 0. Find the inductor current iL as function
of time.
Appendix. Solutions to Exercise Problems 301

Fig. P16.1

Solution
We only know how to solve a circuit that contain only one inductor and one resistor.
Fig. P16.1 has two resistors!

To reduce this problem to a problem that we are able to solve, similar to Problem
15.1, we treat the capacitor as the circuit load and find the Thévenin equivalent
circuit of Fig. P16.1, obtaining Fig. S16.1.

Fig. S16.1

At the initial time t ¼ 0,

vC ð0 Þ ¼ 0:

Since the capacitor’s voltage cannot suddenly change,

vC ð0þ Þ ¼ 0:

At the final time t ¼ 1, the capacitor can be treated an open circuit. Thus,

vC ð1Þ ¼ 1 V:

The time constant is calculated as

τ ¼ RC ¼ ð500 ΩÞð1 μFÞ ¼ 0:5 ms:

Using the general mathematical expression of the capacitor voltage


302 Appendix. Solutions to Exercise Problems

vC ðt Þ ¼ vC ð1Þ þ ½vC ð0Þ  vC ð1Þet=τ , for t  0,

we have

vC ðt Þ ¼ 1  et=ð0:5msÞ V, for t  0:

Problem 16.2 We use the same circuit as in Problem 16.1. We assume that the
switch has been closed for a long time. The switch opens at t ¼ 0. Find the
capacitor’s voltage vC as function of time.

Fig. P16.2

Solution
After the switch opens, only one resistor on the right is effective in the circuit, and
the Thénenin equivalent circuit in Fig. S16.1 is no longer valid.

The initial condition of this problem is the final condition in Problem 16.1.
Using the result of Problem 16.1, at the initial time t ¼ 0,

vC ð0 Þ ¼ 1 V:

Since the capacitor’s voltage cannot suddenly change,

vC ð0þ Þ ¼ 1 V:

At the final time t ¼ 1, the capacitor is discharged to 0. Thus,

vC ð1Þ ¼ 0:

The time constant is calculated as

τ ¼ RC ¼ ð1 kΩÞð1 μFÞ ¼ 1 ms:

Using the general mathematical expression of the capacitor voltage

vC ðt Þ ¼ vC ð1Þ þ ½vC ð0Þ  vC ð1Þet=τ , for t  0,

we have
Appendix. Solutions to Exercise Problems 303

vC ðt Þ ¼ 1  et V, for t  0:

Problem 16.3 The switch in the circuit in Fig. P16.3 has been closed for a long time
before opening at t ¼ 0. Find the capacitor’s voltage vC and the capacitor’s current iC
for t  0.

Solution
When the switch is closed for a long time, the capacitor is an open circuit and the

Fig. P16.3

capacitor’s voltage vC can be solved by Ohm’s law for the 1 Ω resistor.

vC ¼ ð1 ΩÞð1  2vC Þ,

3vC ¼ 1,
1
vC ¼ V:
3
This value is the voltage on the capacitor when the switch has been closed for a
long time. At t ¼ 0, the switch opens, we have

1
v C ð 0þ Þ ¼ v C ð 0 Þ ¼ V:
3
After time t ¼ 0, the switch opens, and Fig. P17.3 becomes Fig. S16.3a.

Fig. S16.3a
304 Appendix. Solutions to Exercise Problems

We do not know how to solve a circuit with an inductor and a dependent source.
Fortunately, our old friend “Thévenin equivalent circuit” is able to help.
This dependent source is the same as that in Problem 15.3, Its Thévenin equiva-
lent circuit for the dependent source is the same as that derived in Problem 15.3 and
is nothing but a resistor of 0.5 Ω. Using the Thévenin equivalent circuit, we reach
Fig. S16.3b.

Fig. S16.3b

There is no source in Fig. S16.3b, and the capacitor’s voltage will eventually
diminish to 0. Thus,

vC ð1Þ ¼ 0:

The time constant is calculated from Fig. S16.3b as

τ ¼ RC ¼ ð0:5 ΩÞð1 FÞ ¼ 0:5 s:

Using the general mathematical expression of the capacitor voltage

vC ðt Þ ¼ vC ð1Þ þ ½vC ð0Þ  vC ð1Þet=τ , for t  0,

we have

1
vC ðt Þ ¼ et=0:5 V, for t  0:
3
Finally, we find capacitor’s current iC for t  0 according to the relationship

dvC ðt Þ
i C ðt Þ ¼ C :
dt
Since C ¼ 1 F, and we have

1 det=0:5 2 t=0:5
i C ðt Þ ¼ ¼ e A, for t  0:
3 dt 3
Appendix. Solutions to Exercise Problems 305

Chapter 17. Analysis of a Circuit by Solving Differential Equations

Problem 17.1 Set up a node equation for the circuit in Fig. P17.1. Then express the
equation in terms of iL.

Fig. P17.1

Solution
The node equation is

vL
¼ ð1 AÞ  ð2vL Þ  ðiL Þ:

For the inductor, we have

diL di
vL ¼ L ¼ ð1 H Þ L ,
dt dt
and the node equation becomes a differential equation

diL di
¼ 1  2 L  iL ,
dt dt
diL
3 þ iL  1 ¼ 0:
dt

Problem 17.2 Set up a differential equation for i1 + i2.


306 Appendix. Solutions to Exercise Problems

Fig. P17.2

Solution
For indictors, the voltage and current follow the relationships

di1 ðt Þ di ðt Þ
v ð t Þ ¼ L1 and vðt Þ ¼ L2 2 ,
dt dt
1 di ðt Þ 1 di ðt Þ
vð t Þ ¼ 1 and vð t Þ ¼ 2 ,
L1 dt L2 dt
 
1 1 d½i ðt Þ þ i2 ðt Þ
þ vð t Þ ¼ 1 :
L1 L2 dt

The KCL equation for the circuit is

vð t Þ
i1 ðt Þ þ i2 ðt Þ þ ¼ 0,
R
L d½i1 ðt Þ þ i2 ðt Þ
i1 ðt Þ þ i2 ðt Þ þ ¼ 0,
R dt
L diðt Þ
þ iðt Þ ¼ 0,
R dt
where

iðt Þ ¼ i1 ðt Þ þ i2 ðt Þ,
1 1 1
¼ þ :
L L1 L2
The differential equation

L d½i1 ðt Þ þ i2 ðt Þ
i1 ðt Þ þ i2 ðt Þ þ ¼0
R dt
is solvable by solving
Appendix. Solutions to Exercise Problems 307

L diðt Þ
þ iðt Þ ¼ 0,
R dt
whose solution is

iðt Þ ¼ ið1Þ þ ½ið0Þ  ið1ÞeRt=L , for t  0:

This circuit does not contain any sources, the combined inductor current i(t) will
eventually diminish to 0. That is,

iðt Þ ¼ ið0ÞeRt=L , for t  0:

This result i(1) ¼ 0 does not imply i1(1) ¼ i2(1) ¼ 0. It only implies

i1 ð1Þ ¼ i2 ð1Þ ¼ constant:

What is this constant? In fact, any constant will satisfy the differential equation.
We need to find the constant satisfy the energy conservation.
The energy stored in an inductor is

1
w ¼ Li2 :
2
The initial energy of the system is determined by the initial currents in the
inductors. The initial energy is, therefore,

1 1
w0 ¼ L1 i21 ð0Þ þ L2 i22 ð0Þ:
2 2
The power consumption of the resistor is

pR ¼ i2 R ¼ Ri2 ð0Þe2Rt=L , for t  0:

The total energy consumed by the resistor is

Z1 Z1
L
wR ¼ pR dt ¼ Ri2 ð0Þe2Rt=L dt ¼ i2 ð0Þ:
2
0 0

The final energy stored in the inductors is

1 1 i 2 ð 1Þ
w1 ¼ L1 i21 ð1Þ þ L2 i22 ð1Þ ¼ 1 ðL1 þ L2 Þ:
2 2 2
Energy conservation demands

w1 ¼ w0  wR ,
308 Appendix. Solutions to Exercise Problems

i21 ð1Þ 1 1 L
ðL1 þ L2 Þ ¼ L1 i21 ð0Þ þ L2 i22 ð0Þ  i2 ð0Þ,
2 2 2 2
L1 i21 ð0Þ þ L2 i22 ð0Þ  Li2 ð0Þ
i21 ð1Þ ¼ :
L1 þ L2
In order to get some intuition, let us consider two numerical examples:

Example 1: L1 ¼ L2 ¼ 1 H and i1(0) ¼ i2(0) ¼ 1 A, we have i1(1) ¼  i2(1) ¼ 0.


Example 2: L1 ¼ L2 ¼ 1 H, i1(0) ¼ 1 A and i2(0) ¼ 0, we have
i1(1) ¼  i2(1) ¼ 0.5 A.

This problem assumes ideal inductors, where the conductor is perfect with zero
resistance. For an everyday inductor, we do not have i(1) ¼ 0.5 A. The inductor
current will eventually diminish and get i(1) ¼ 0. The energy stored in the inductor
is in the form of magnetic field.
Nowadays, the medical MRI machine uses liquid helium to create a
superconducting environment so that the superconductor coil can maintain a strong
magnetic field for a long time.

Problem 17.3 Set up a differential equation for v1  v2.

Fig. P17.3

Solution
The KVL equation for the circuit is

v1 ðt Þ  v2 ðt Þ  R  iðt Þ ¼ 0:

For capacitors, the voltage and current follow the relationships

dv2 ðt Þ
iðt Þ ¼ C 2 ,
dt
Appendix. Solutions to Exercise Problems 309

dv1 ðt Þ
iðt Þ ¼ C 1 :
dt
We have

i ð t Þ i ð t Þ d½ v 1 ð t Þ  v 2 ð t Þ 
  ¼ :
C1 C2 dt
Let

1 1 1
¼ þ and vðt Þ ¼ v1 ðt Þ  v2 ðt Þ:
C C1 C1
We have

iðt Þ iðt Þ d½v1 ðt Þ  v2 ðt Þ


  ¼ ,
C1 C1 dt
d½v1 ðt Þ  v2 ðt Þ
iðt Þ ¼ C ,
dt
d½v1 ðt Þ  v2 ðt Þ
½v1 ðt Þ  v2 ðt Þ þ RC ¼ 0,
dt
dvðt Þ
vðt Þ þ RC ¼ 0:
dt
The differential equation

dvðt Þ
vðt Þ þ RC ¼0
dt
has a general solution with τ ¼ RC

vðt Þ ¼ vð1Þ þ ½vð0Þ  vð1Þet=τ , for t  0:

Our circuit does not contain any sources, the combined capacitor voltage v(t) will
eventually diminish to 0. That is,

vðt Þ ¼ vð0Þet=τ , for t  0:

This result v(1) ¼ 0 does not imply v1(1) ¼ v2(1) ¼ 0. It only implies

v1 ð1Þ ¼ v2 ð1Þ ¼ constant:

What is this constant? In fact, any constant will satisfy the differential equation.
We need to find the constant satisfy the energy conservation.
The energy stored in a capacitor is
310 Appendix. Solutions to Exercise Problems

1
w ¼ Cv2 :
2
The initial energy of the system is determined by the initial voltages in the
capacitors. The initial energy is, therefore,

1 1
w0 ¼ C1 v21 ð0Þ þ C 2 v22 ð0Þ:
2 2
The power consumption of the resistor is

v2 v2 ð0Þ 2t=τ
pR ¼ ¼ e , for t  0:
R R
The total energy consumed by the resistor is

Z1 Z1
v2 ð0Þ 2t=τ τ 2 C
wR ¼ pR dt ¼ e dt ¼ v ð0Þ ¼ v2 ð0Þ:
R 2R 2
0 0

The final energy stored in the inductors is

1 1 v2 ð 1Þ
w1 ¼ C1 v21 ð1Þ þ C 2 v22 ð1Þ ¼ 1 ðC 1 þ C2 Þ:
2 2 2
Energy conservation demands

w1 ¼ w0  wR ,

v21 ð1Þ 1 1 C
ðC 1 þ C 2 Þ ¼ C1 v21 ð0Þ þ C2 v22 ð0Þ  v2 ð0Þ,
2 2 2 2
C1 v21 ð0Þ þ C 2 v22 ð0Þ  Cv2 ð0Þ
v21 ð1Þ ¼ :
C1 þ C2
In order to get some intuition, let us consider two numerical examples:

Example 1: C1 ¼ C2 ¼ 1 F and v1(0) ¼ v2(0) ¼ 1 V, we have v1(1) ¼ v2(1) ¼ 0.


Example 2: C1 ¼ C2 ¼ 1 F, v1(0) ¼ 1 V and v2(0) ¼ 0, we have
v1(1) ¼ v2(1) ¼ 0.5 V.

This problem assumes ideal capacitors, where the dielectric between the two
metal plates is perfect insulator. For an everyday capacitor, we do not have v
(1) ¼ 0.5 V. The capacitor will be eventually discharged and get v(1) ¼ 0. The
energy stored in the capacitor can be expressed in voltage (v) or in charge (Q) as

Cv2 Q2
w¼ ¼ :
2 2C
Appendix. Solutions to Exercise Problems 311

Nowadays, the supercapacitors can store electric charges and used as batteries,
but their self-discharge rate is significantly faster than rechargeable batteries.

Chapter 18. First-Order Circuits

Problem 18.1 The input of an RC circuit is a periodic square pulse sequence. The
period is 2 T. The time constant of the RC circuit is τ. The output signal is the
capacitor voltage vC. Find the output signal’s maximum value vmax and the minimum
value vmin.

Fig. P18.1
312 Appendix. Solutions to Exercise Problems

Solution
This is a sequential switching first-order system. We will use this general solution
twice.

vC ðt Þ ¼ vC ð1Þ þ ½vC ð0Þ  vC ð1Þet=τ , for t  0,

where t ¼ 0 is understood as the initial time.


For the charging period,

vC ð0Þ ¼ vmin ,

vC ð1Þ ¼ 1,

vC ðT Þ ¼ vmax ,

vC ðT Þ ¼ 1 þ ½vmin  1eT=τ ¼ vmax :

For the discharging period,

vC ð0Þ ¼ vmax ,

vC ð1Þ ¼ 0,

vC ðT Þ ¼ vmin ,

vC ðT Þ ¼ vmax eT=τ ¼ vmin :

From
(
1 þ ½vmin  1eT=τ ¼ vmax
vmax eT=τ ¼ vmin

we obtain
8
>
> 1  eT=τ
< vmax ¼ ,
1  e2T=τ
> T=τ
: vmin ¼ e  1 :
>
e 2T=τ 1
If the time constant τ is much longer than T, we have

vmax  0:5,
vmin  0:5:

This circuit can be used to smooth the input signal. As we will find out later in this
book, this circuit is a lowpass filter.
Appendix. Solutions to Exercise Problems 313

Problem 18.2 A student tries to solve a problem in his own way, and he does not
get the correct answer. Please help him to find the error. In the problem, the switch
has been closed for a long time. The switch opens at t ¼ 0. Find the capacitor’s
voltage iC as function of time.

Fig. P18.2

The student’s solution:


After the switch has been closed for a long time, the capacitor acts like an open
circuit. Therefore,

iC ð0Þ ¼ 0:

At t ¼ 0, the switch opens. Now the circuit does not have any source, and the
capacitor will eventually discharge to 0. Thus,

iC ð1Þ ¼ 0:

Recall the general solution

iðt Þ ¼ ið1Þ þ ½ið0Þ  ið1Þet=τ , for t  0:

The student’s solution is

iC ðt Þ ¼ 0, for t  0:

Solution
For the RC circuit, we need to solve the capacitor’s voltage first. The capacitor’s
voltage has a general solution
314 Appendix. Solutions to Exercise Problems

vC ðt Þ ¼ vC ð1Þ þ ½vC ð0Þ  vC ð1Þet=τ , for t  0:

After that capacitor’s voltage vC(t) is found, we can use vC(t) to find other
unknowns. For example, if we want to find the capacitor’s current iC(t), we need
to use the formula

dvC ðt Þ
i C ðt Þ ¼ C :
dt
If we want to find the voltage across the resistor, we use vC(t) and KVL. The
voltage across a capacitor cannot change suddenly; however, the current through a
capacitor can change suddenly. In an RC circuit, the voltage across the capacitor
determines the behavior of the entire circuit.
Likewise, for the RL circuit, we need to solve the inductor’s current first. The
inductor’s current has a general solution

iL ðt Þ ¼ iL ð1Þ þ ½iL ð0Þ  iL ð1Þet=τ , for t  0,

After that inductor’s current iL(t) is found, we can use iL(t) to find other
unknowns. For example, if we want to find the inductor’s voltage vL(t), we need to
use the formula

diL ðt Þ
vL ð t Þ ¼ L :
dt
If we want to find the current through the resistor, we use iL(t) and KCL. The
current through an inductor cannot change suddenly; however, the voltage across an
inductor can change suddenly. In an RL circuit, the current through the inductor
determines the behavior of the entire circuit.

Chapter 19. Sinusoidal Steady State

Problem 19.1 Express the following signals in the phasor form:

(a) 5 cos (100t)


(b) 5 sin ð100t Þ
 
(c) 5 cos 100t þ 45
 
(d) 5 sin 100t þ 45
(e) 2 cos ðωt Þ
(f) 2 cos ðωt Þ  3 cos ð2ωt Þ
(g) 10
(h) 2t2 cos (ωt)
Appendix. Solutions to Exercise Problems 315

Solution

(a) 5 ∠ 0
 
(b) 5 sin (100t) ¼ 5 cos (100t  90 ), so, the phasor form is 5 ∠  90 .

(c) 5∠45
   
(d) 5 sin (100t + 45 )¼ 5 cos (100t + 45  90 ), so, the phasor form is 5 ∠  45 .
(e) 5
(f) 2 cos (ωt)  3 cos (2ωt). This expression has two different frequencies. Cannot
use phasor forms.
(g) This is a dc signal. Do not use sinusoidal analysis.
(h) Not a sinusoidal signal due to the 2t2 time-varying amplitude.

Problem 19.2 Express the transfer function in the phasor form. The input is vin and
the output is vC.

Fig. P19.2

Solution
This is a voltage divider, and the capacitor can be treated as a “resistor” with
impedance (similar to “resistance”)

1 1 
Z¼ ¼ ∠  90 :
jωC ωC
The transfer function is the ratio of output over input, which is, in the phasor
form,

Z
H ð ωÞ ¼ ,
RþZ
1
jωC
¼ ,
R þ jωC
1
316 Appendix. Solutions to Exercise Problems

1
¼ ,
jωCR þ 1
1
¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ,
2
ðωCRÞ þ 1∠ tan 1 ðωCRÞ

1
¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ∠  tan 1 ðωCRÞ:
ðωCRÞ2 þ 1

You may wonder why we are interested in a transfer function. A transfer function
comes in handy if you know the input and want to find the output.
Here is a numerical solution.
Let R ¼ 1 kΩ, C ¼ 1 μF, and vin(t) ¼ 10 cos (1000t). Thus, ω ¼ 1000 and the
input phasor is just

10∠0 :

In fact, when we convert the time-domain signal vin(t) ¼ 10 cos (1000t) to the
phasor
    
10∠0 ¼ 10 cos 0 þ j10 sin 0 ,

we have done two things. First, we discard the frequency ω ¼ 1000. Second, we add
an imaginary term.
The output phasor in the product of the input phasor and the transfer function
2 3
 6
  1 7
10∠0  4qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ∠  tan 1 ðωCRÞ5
ðωCRÞ2 þ 1
 
  1
¼ 10∠0  pffiffiffi ∠  tan 1 ð1Þ ,
2
 
  1 
¼ 10∠0  pffiffiffi ∠  45 ,
2
10 
¼ pffiffiffi ∠  45 :
2
You need to remind yourself that a phasor is just a complex number, which can be
in the Cartesian form

a þ jb

or in the polar form (exponential form)


Appendix. Solutions to Exercise Problems 317

pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi   pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 b j tan 1 ðbaÞ
a þ b ∠ tan
2 2
¼ a2 þ b2 e :
a

The product of two phasors is the product of two complex numbers. In the polar
form, the magnitude is the product of the two magnitudes and the phase is the sum of
the phases.
It is easier to use the Cartesian form to add and subtract phasors. It is easier to use
the polar form (or exponential form) to multiply and divide phasors.
In the regular time-domain expression, the output signal is

10  
vC ðt Þ ¼ pffiffiffi cos 1000t  45 :
2
We must remember two things when converting a phasor back to the time-domain
signal. First, we take real part (i.e., keeping the cosine function) and discard the
imaginary part (i.e., removing the sine function). Second, insert the frequency
ω ¼ 1000 in the cosine function as ωt. This frequency ω must be the same frequency
ω in the input signal.

Chapter 20. Function Generators and Oscilloscopes

Problem 20.1 In Fig. P20.1, an oscilloscope is directly connected to a signal


generator to verify the signal generated. We set the peak-to-peak voltage of a
sinewave to be 10 V. However, the oscilloscope shows a 20-V peak-to-peak. Is
there anything wrong?

Fig. P20.1
318 Appendix. Solutions to Exercise Problems

Solution
Nothing is wrong. This issue has already been discussed in text. We will discuss it
again here.

The default setting for most function generators is to display the desired voltage
as though terminated into a 50-Ohm load. When a high impedance device, such as an
oscilloscope is used to measure the output of the function generator, the waveform
appears to be twice the voltage set on the display of the oscilloscope.

Fig. S20.1

The Thévenin equivalent circuit for a typical signal generator is shown in


Fig. S20.1. The value of vout is displayed on the generator. If load is 50 Ω, the actual
vin is twice as large as vout. Even if the load is large, the generator’s display ignores
the actual vout and displays a value that is half of the value of vin.
The input impedance of an oscilloscope is in the order of 1 MΩ; therefore, the
oscilloscope shows almost a correct amplitude of vin.

Problem 20.2 How to use an oscilloscope to estimate the time constant of a first-
order circuit?

Solution
If your first-order circuit is an RC circuit, you can use an oscilloscope to measure the
voltage across the capacitor.

If your first-order circuit is an RL circuit, do not use an oscilloscope to measure


the voltage across the inductor because this voltage is almost zero. In this case, find a
resistor in the circuit and measure the voltage drop across this resistor.
A portion of your signal is described by an exponential function

½vð0Þ  vð1Þet=τ :

This may be a charging or discharging trend, as shown in Fig. S20.2.


Appendix. Solutions to Exercise Problems 319

Fig. S20.2

Let t ¼ τ, and eτ/τ ¼ e1 ¼ 0.368. Find the value that is

v1 ¼ 0:368  jvð1Þ  vð0Þj:

Adjust the oscilloscope’s vertical and time scaling knobs so that you see a
charging or discharging period nice and big.
Let us assume that you are looking at the charging period.
When you hit the cursor button on the oscilloscope, a menu should come up on
the screen saying “Cursors.” Set the time position of the cursor at the point the
charging begins. This is your t1. Then move the time position of the cursor to
position the curve, that is v1, above the starting position. This is your t2. Your time
constant is

τ ¼ t2  t1:
320 Appendix. Solutions to Exercise Problems

Chapter 21. Mutual Inductance and Transformers

Problem 21.1 An ideal transformer has 1000 turns in its primary coil and 100 turns
in its secondary coil. Determine whether the following statements are true.

(a)This is a 10:1 transformer.


(b)This is a 1:10 transformer.
(c)This is a 1:0.1 transformer.
(d)This is a 0.1:1 transformer.
(e)This is a step-up transformer.
(f)
This is a step-down transformer.
(g)If the primary voltage is 10 V, the secondary voltage is 100 V.
(h)If the primary voltage is 10 V, the secondary voltage is 1 V.
(i)
If the primary current is 10 A, the secondary current is 100 A.
(j)
If the primary current is 10 A, the secondary current is 1 A.
(k)The transformer consumes power.
(l)
The transformer only works for a DC source input.
(m) The transformer only works for an AC source input.
(n)The frequency on secondary side is 10 times higher than the frequency on the
primary side.
(o) The frequency on secondary side is 10 times lower than the frequency on the
primary side.
(p) If the secondary side has a load of 100 Ω, the reflected impedance on the
primary side is 1000 Ω.
(q) If the secondary side has a load of 100 Ω, the reflected impedance on the
primary side is 10,000 Ω.

Solution
(a) This is a 10:1 transformer. This is true.
(b) This is a 1:10 transformer. This is false.
(c) This is a 1:0.1 transformer. This is true.
(d) This is a 0.1:1 transformer. This is false.
(e) This is a step-up transformer. This is false.
(f) This is a step-down transformer. This is true.
(g) If the primary voltage is 10 V, the secondary voltage is 100 V. This is false.
(h) If the primary voltage is 10 V, the secondary voltage is 1 V. This is true.
(i) If the primary current is 10 A, the secondary current is 100 A. This is true.
(j) If the primary current is 10 A, the secondary current is 1 A. This is false.
(k) The transformer consumes power. This is false.
(l) The transformer only works for a DC source input. This is false.
(m) The transformer only works for an AC source input. This is true.
(n) The frequency on secondary side is 10 times higher than the frequency on the
primary side. This is false. The frequency does not change.
(o) The frequency on secondary side is 10 times lower than the frequency on the
primary side. This is false.
Appendix. Solutions to Exercise Problems 321

(p) If the secondary side has a load of 100 Ω, the reflected impedance on the
primary side is 1000 Ω. This is false.
(q) If the secondary side has a load of 100 Ω, the reflected impedance on the
primary side is 10,000 Ω. This is true.

Problem 21.2 This problem is about the dot notation and convention in a trans-
former. Express the induced voltages for each case.

(a)

Fig. P21.2a

(b)

Fig. P21.2b

(c)

Fig. P21.2c

(d)

Fig. P21.2d
322 Appendix. Solutions to Exercise Problems

(e)

Fig. P21.2e

(f)

Fig. P21.2f

(g)

Fig. P21.2g

(h)

Fig. P21.2h

Solution
(a) v2 ¼ M didt1 and v1 ¼ M didt2 :
(b) v2 ¼ M didt1 and v1 ¼ M didt2 :
Appendix. Solutions to Exercise Problems 323

(c) v2 ¼ M didt1 and v1 ¼ M didt2 :


(d) v2 ¼ M didt1 and v1 ¼ M didt2 :
(e) v2 ¼ M didt1 and v1 ¼ M didt2 :
(f) v2 ¼ M didt1 and v1 ¼ M didt2 :
(g) v2 ¼ M didt1 and v1 ¼ M didt2 :
(h) v2 ¼ M didt1 and v1 ¼ M didt2 :

Chapter 22. Fourier Series

Problem 22.1 Match a Fourier series with a periodic function with ω0 ¼ 2π/T.

X1
f 1 ð t Þ ¼ a0 þ a
n¼1 n
cos ðnω0 t Þ
X1
f 2 ðt Þ ¼ a
n¼1 n
cos ðnω0 t Þ
X1
f 3 ð t Þ ¼ a0 þ b
n¼1 n
sin ðnω0 t Þ
X1
f 4 ðt Þ ¼ b
n¼1 n
sin ðnω0 t Þ
X1
f 5 ðt Þ ¼ n ¼ 1 ½an cos ðnω0 t Þ þ bn sin ðnω0 t Þ
n ¼ odd
X1
f 6 ðt Þ ¼ n ¼ 1 an cos ðnω0 t Þ
n ¼ odd
X1
f 7 ðt Þ ¼ n ¼ 1 bn sin ðnω0 t Þ
n ¼ odd
X1
f 8 ð t Þ ¼ a0 þ n¼1
½an cos ðnω0 t Þ þ bn sin ðnω0 t Þ
X1
f 9 ðt Þ ¼ n¼1
½an cos ðnω0 t Þ þ bn sin ðnω0 t Þ

(a)

Fig. P22.1a
324 Appendix. Solutions to Exercise Problems

(b)

Fig. P22.1b

(c)

Fig. P22.1c

(d)

Fig. P22.1d

(e)

Fig. P22.1e

(f)

Fig. P22.1f
Appendix. Solutions to Exercise Problems 325

(g)

Fig. P22.1g

(h)

Fig. P22.1h

(i)

Fig. P22.1i

Solution P
(a) f 6 ðt Þ ¼ 1n ¼ 1 an cos ðnω0 t Þ

P1n ¼ odd
(b) f 7 ðt Þ ¼ n ¼ 1 bn sin ðnω0 t Þ
P n ¼ odd
(c) f 5 ðt Þ ¼ 1n ¼ 1 ½an cos ðnω0 t Þ þ bn sin ðnω0 t Þ

P n ¼ odd
(d) f 4 ðt Þ ¼ 1 bn sin ðnω0 t Þ
n¼1P
(e) f 3 ð t Þ ¼ a0 þ 1 b sin ðnω0 t Þ
Pn¼1 n
(f) f 1 ð t Þ ¼ a0 þ 1 n¼1 an cos ðnω0 t Þ
P
(g) f 2 ðt Þ ¼ 1 a cos ðnω0 t Þ
Pn¼1 n
(h) f 9 ðt Þ ¼ 1 n¼1Pn cos ðnω0 t Þ þ bn sin ðnω0 t Þ
½a
(i) f 8 ð t Þ ¼ a0 þ 1 n¼1 ½an cos ðnω0 t Þ þ bn sin ðnω0 t Þ:

Problem 22.2 Show that the Fourier series has an equivalent exponential form

X1
jnω0 t
f ðt Þ ¼ ce
n¼1 n
:
326 Appendix. Solutions to Exercise Problems

Solution
Using Euler’s formula

ejnω0 t ¼ cos ðnω0 t Þ þ j sin ðnω0 t Þ,

We have

ejnω0 t þ ejnω0 t
cos ðnω0 t Þ ¼ ,
2
ejnω0 t  ejnω0 t
sin ðnω0 t Þ ¼ :
2j

The Fourier series


X1
f ð t Þ ¼ a0 þ n¼1
½an cos ðnω0 t Þ þ bn sin ðnω0 t Þ

can be rewritten as

X1 ejnω0 t þ ejnω0 t ejnω0 t  ejnω0 t


f ð t Þ ¼ a0 þ n¼1
an þ bn ,
2 2j
X1 an  jbn jnω0 t an þ jbn jnω0 t
f ð t Þ ¼ a0 þ n¼1
e þ e ,
2 2
X1
f ð t Þ ¼ c0 þ c ejnω0 t þ cn ejnω0 t :
n¼1 n

where

an  bn
cn ¼ ,
2
c 0 ¼ a0 :

Chapter 23. Laplace Transform in Circuit Analysis

Problem 23.1 Solve the following differential equation using the Laplace transform
method.

x} ðt Þ þ 4x0 ðt Þ þ 3xðt Þ ¼ 5

with initial conditions x0(0) ¼ 1 and x(0) ¼ 2.


Appendix. Solutions to Exercise Problems 327

Solution
Taking the Laplace transform term by term according to the Laplace transform table,
we have

5
s2 X ðsÞ  sxð0Þ  x0 ð0Þ þ 4sX ðsÞ  4xð0Þ þ 3X ðsÞ ¼ ,
s
5
s2 X ðsÞ  2s  1 þ 4sX ðsÞ  8 þ 3X ðsÞ ¼ ,
s

s þ 9 þ 2s
5
5 þ 9s þ 2s2 k k k
X ðsÞ ¼ ¼ ¼ 1þ 2 þ 3 :
s2 þ 4s þ 3 sðs þ 1Þðs þ 3Þ s sþ1 sþ3

We can use the cover-up method to find the partial fraction expansion.

5 þ 9ð0Þ þ 2ð0Þ2 5
k1 ¼ ¼ ,
ð 0 þ 1Þ ð 0 þ 3Þ 3

5 þ 9ð1Þ þ 2ð1Þ2
k2 ¼ ¼ 1,
ð1Þð1 þ 3Þ

5 þ 9ð3Þ þ 2ð3Þ2 2
k3 ¼ ¼ :
ð3Þð3 þ 1Þ 3

The Laplace transform of the solution is

5=3 1 2=3
X ðsÞ ¼ þ þ :
s sþ1 sþ3
The solution can be obtained by finding the inverse Laplace transform (using the
Table of Laplace Transform Pairs):

5 2
xð t Þ ¼ þ et  e3t , for t  0:
3 3

Problem 23.2 Use the Laplace transform method to solve for the circuit in
Fig. P23.2. The initial voltage in the capacitor is 3 V.

Fig. P23.2
328 Appendix. Solutions to Exercise Problems

Solution
We first need to convert the time-domain circuit in Fig. P23.2 to its corresponding
Laplace-domain circuit in Fig. S23.2a. The initial condition is converted into a
voltage source or a current source.

We know that if the initial condition is zero, the time-domain relationship for the
capacitor

dvðt Þ
i ðt Þ ¼ C
dt
has a Laplace-domain counterpart

I ðsÞ ¼ CsV ðsÞ:

According to the Laplace transform table, if the initial condition is not zero, the
Laplace-domain counterpart is

v ð 0Þ
I ðsÞ ¼ C ½sV ðsÞ  vð0Þ ¼ CsVðsÞ  Cvð0Þ ¼ Cs V ðsÞ  :
s

There are two ways to handle this initial condition. The first way is to use the
relationship

I ðsÞ ¼ CsV ðsÞ  Cvð0Þ

and treat the initial condition as a current source of value

Cvð0Þ ¼ 3 μ:

The second way is to use the relationship

I ðsÞ vð0Þ
V ðsÞ ¼ þ
Cs s
and treat the initial condition as a voltage source of value

v ð 0Þ 3
¼ :
s s
Therefore, we can have two Laplace-main circuits, as shown in Fig. S23.2a and
Fig. S23.2b, respectively.
Appendix. Solutions to Exercise Problems 329

Fig. S23.2a

Fig. S23.2b

Let us first solve the circuit in Fig. S24.1. The node equation is

V  2s V V
þ 1 þ ¼ 3 μ,
1k sμ
1 k

3 μ þ s 12 k 3s þ 2000 k k2
V¼ ¼ ¼ 1þ :
1kþsμ
2 sð2000 þ sÞ s s þ 2000

Using the cover-up method to find k1 and k2,

3ð0Þ þ 2000
k1 ¼ ¼ 1,
2000 þ 0
3ð2000Þ þ 2000
k2 ¼ ¼ 2:
ð2000Þ

Therefore,
330 Appendix. Solutions to Exercise Problems

1 2
V¼ þ
s s þ 2000
and

vðt Þ ¼ 1 þ 2e2000t V, for t  0:

Now let us solve the circuit in Fig. S23.2b. The node equation is

V  2s V  3s V
þ 1 þ ¼ 0,
1k sμ
1 k

sþ3
2000
3s þ 2000
V¼ ¼ :
2000 þ s sð2000 þ sÞ

This expression is exactly the same as that for Fig. S23.2a. Therefore, we have the
same solution

vðt Þ ¼ 1 þ 2e2000t V, for t  0:

Problem 23.3 Use the Laplace transform method to solve for the circuit in
Fig. P23.3. The initial current in the inductor is 3 A.

Fig. P23.3

Solution
We first need to convert the time-domain circuit in Fig. P23.3 to its corresponding
Laplace-domain circuit in Fig. S23.3a. The initial condition is converted into a
voltage source or a current source.

We know that if the initial condition is zero, the time-domain relationship for the
inductor
Appendix. Solutions to Exercise Problems 331

diðt Þ
vð t Þ ¼ L
dt
has a Laplace-domain counterpart

V ðsÞ ¼ LsI ðsÞ:

According to the Laplace transform table, if the initial condition is not zero, the
Laplace-domain counterpart is

ið0Þ
V ðsÞ ¼ L½sI ðsÞ  ið0Þ ¼ LsI ðsÞ  Lið0Þ ¼ Ls I ðsÞ  :
s

There are two ways to handle this initial condition. The first way is to use the
relationship

V ðsÞ ¼ LsI ðsÞ  Lið0Þ

and treat the initial condition as a voltage source of value

Lið0Þ ¼ 3:

The second way is to use the relationship

V ðsÞ ið0Þ
I ðsÞ ¼ þ
Ls s
and treat the initial condition as a current source of value

i ð 0Þ 3
¼ :
s s
Therefore, we can have two Laplace-main circuits, as shown in Fig. S23.3a and
Fig. S23.3b, respectively.

Fig. S23.3a
332 Appendix. Solutions to Exercise Problems

Fig. S23.3b

Let us first solve the circuit in Fig. S23.3a. The node equation is

V V 1 3
þ ¼  2V  ,
1 s s s
2
V ¼  :
3 s þ 13

The current through the inductor is i(t) and its Laplace transform is

V 2 2 2
I¼ ¼  ¼ þ :
s 3s s þ 3
1 s s þ 13

After taking the inverse Laplace transform using the Table of Laplace Transform
Pairs,

iðt Þ ¼ 2 þ 2e3t A, for t  0:


1

Now let us solve the circuit in Fig. S23.3b. The node equation is

V V þ3 1
þ ¼  2V,
1 s s
2
V ¼  :
3 s þ 13

This expression is exactly the same as that for Fig. S23.3a. Therefore, we have the
same solution.
Appendix. Solutions to Exercise Problems 333

Chapter 24. Fourier Transform in Circuit Analysis

Problem 24.1 Use the Fourier transform method to solve for the circuit in
Fig. P24.1. The initial voltage in the capacitor is 3 V.

Fig. P24.1

Solution
This problem is the same as Problem 23.2, in which the Laplace transform method is
used. The difficult part of solving this problem using the Fourier transform is that the
Fourier transform method does not have an explicit way to handle the initial
conditions.

In order to accommodating an initial condition, we need to introduce a source


with a step function as follows.
In Problem 23.2, the initial condition of the capacitor is treated as a voltage source
of value in the Laplace domain

v ð 0Þ 3
¼ :
s s
Converting this source into the time domain gives the time-domain source

vð0Þuðt Þ ¼ 3 uðt Þ,

where u(t) is the unit step function. The Fourier transform of the unit step function
can be found in the Table of Fourier transform Pairs as

1
πδðωÞ þ :

We can redraw the circuit by introducing the step function sources in Fig. S24.1.

We now set up a node equation for the circuit in Fig. S24.1 in the Fourier domain
as follows.
334 Appendix. Solutions to Exercise Problems

Fig. S24.1

h i h i
V  2 πδðωÞ þ jω
1
V  3 πδðωÞ þ jω
1
V
þ 1
þ ¼ 0,
1k jωμ
1k

1 1
V  2 πδðωÞ þ þ jωmV  3jωm πδðωÞ þ þ V ¼ 0,
jω jω
h i h i
2 πδðωÞ þ jω
1
þ 3jωm πδðωÞ þ jω 1
V¼ ,
2 þ jωm
h i h i
2000 πδðωÞ þ jω 1
þ 3jω πδðωÞ þ jω1
V¼ ,
2000 þ jω

k1 1
V¼ þ k2 πδðωÞ þ :
2000 þ jω jω

It can be verified that k1 ¼ 2 and k2 ¼ 1. After taking the inverse Fourier transform
by using the Table of Fourier Transform Pairs,

vðt Þ ¼ 1 þ 2e2000t V, for t  0:

This answer is the same as that obtained from Problem 23.2, obtained by the
Laplace transform. The voltage is the capacitor voltage, which is normally denoted
by vC, as in Fig. P24.1. This voltage includes the everything inside the dotted box in
Fig. S24.1.
We feel that if a circuit problem involves initial conditions or switch actions, the
Laplace transform method is easier than the Fourier transform method because the
Laplace transform of the unit step function is 1/s, while the Fourier transform of the
unit step function is πδ(ω) + 1/( jω).

Problem 24.2 The input signal is a signum function. Find the inductor current iL.
Appendix. Solutions to Exercise Problems 335

Fig. P24.2

Solution
This problem cannot be solved by the Laplace transform method because the input
signal is not zero when t < 0.

This problem cannot be solved by the phasor method either because the input
signal is not sinusoidal.
This problem can be solved either in the time domain or in the Fourier domain.
We will use the Fourier transform method here.
The Fourier transform of the signum function is

1
:

We notice that this circuit is current divider. Let us set up the current divider
relationship in the Fourier domain as follows.
   
1 1 1 1 1=2 1=2
IC ¼ ¼ ¼  :
jω ð1 þ jωÞ þ 1 jω 2 þ jω jω 2 þ jω

Taking the invers Fourier transform yields

1 1
i C ðt Þ ¼ sgn ðt Þ  e2t uðt Þ A,
2 2
where u(t) is the unit step function.

Problem 24.3 Repeat Problem 24.2 with 10 cos (4t) being the input signal.

Solution
Since the input function is sinusoidal, both the phasor method and the Fourier
method can be used. The Laplace transform method does not work.

The Fourier transform of the input function is


336 Appendix. Solutions to Exercise Problems

10π ½δðω þ 4Þ þ δðω  4Þ:

Let us set up the current divider relationship in the Fourier domain as follows.

1
I C ¼ ð10π ½δðω þ 4Þ þ δðω  4ÞÞ ,
ð1 þ jωÞ þ 1

δðω þ 4Þ δðω  4Þ
¼ 10π þ ,
2 þ jω 2 þ jω

δðω þ 4Þ δðω  4Þ
¼ 10π þ ,
2 þ jð4Þ 2 þ jð4Þ

¼ π ½ð1 þ j2Þδðω þ 4Þ þ ð1  j2Þδðω  4Þ,

¼ π ½ðδðω þ 4Þ þ δðω  4Þ þ 2jπ ½ðδðω þ 4Þ  δðω  4Þ:

Taking the invers Fourier transform yields

iC ðt Þ ¼ cos ð4t Þ  2 sin ð4t Þ A:

Chapter 25. Second-Order Circuits

Problem 25.1 In a series RLC circuit, what does the step response look like when
R ¼ 0? The input is step voltage source, and the output is the voltage across the
capacitor.

Fig. P25.1

Solution
For a step response, it is easier to the Laplace transform method. The Laplace-
domain circuit is shown in Fig. S25.1.
Appendix. Solutions to Exercise Problems 337

Fig. S25.1

This circuit is a voltage divider. We have

  1   1
1 sC 1 LC 1 s
V¼ ¼ ¼  2 :
s sL þ sC
1 s s2 þ LC
1 s s þ LC
1

Taking the inverse Laplace transform yields


 
1
vð t Þ ¼ 1  cos pffiffiffiffiffiffi t uðt Þ:
LC
This function is biased sinewave after t ¼ 0. When R is 0, the series RLC circuit
becomes an oscillator. The effect of the resistance R is to damp the oscillation. A
larger R has a stronger damping effect.

Problem 25.2 In a parallel RLC circuit, what does the step response look like when
R ¼ 1? The input is step current source, and the output is the current through the
inductor.

Fig. P25.2

Solution
For a step response, it is easier to the Laplace transform method. The Laplace-
domain circuit is shown in Fig. S25.2.
338 Appendix. Solutions to Exercise Problems

Fig. S25.2

This circuit is a current divider. We have

  1   1
1 sC 1 LC 1 s
I¼ ¼ ¼  2 :
s sL þ sC
1 s s2 þ LC
1 s s þ LC
1

Taking the inverse Laplace transform yields


 
1
iðt Þ ¼ 1  cos pffiffiffiffiffiffi t uðt Þ:
LC
This function is biased sinewave after t ¼ 0. When R is 1, the parallel RLC
circuit becomes an oscillator. A smaller LC gives a higher oscillating frequency. The
effect of the resistance R is to damp the oscillation. A smaller R has a stronger
damping effect.

Problem 25.3 The circuit in Fig. P25.3 is neither a series nor a parallel circuit. It is
still a second-order RLC circuit. Find the conditions for circuit to be underdamped,
critically damped, and overdamped.

Fig. P25.3
Appendix. Solutions to Exercise Problems 339

Solution
For a step response, it is easier to the Laplace transform method. The Laplace-
domain circuit is shown in Fig. S25.3.

Fig. S25.3

This circuit is a voltage divider. We have


R
  sC   R  
1 RþsC1
1 sC 1 R
V¼  R ¼  R ¼ :
s sL þ sC s sL R þ sC þ sC
1 s s2 RLC þ sL þ R
RþsC
1

In the above equation, the input is 1/s and the transfer function is

1
R LC
H¼ ¼ :
s RLC þ sL þ R s2 þ s RC
2 1
þ LC
1

Letting the denominator of the transfer function be 0, we have the characteristic


equation of the circuit

1 1
s2 þ s þ ¼ 0:
RC LC
The solutions of this characteristic equation are given as
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 RC
1
 1
ðRC Þ2
 LC 4

s1,2 ¼ :
2
The condition for a critically damped circuit is when the two solutions are
identical, that is,

1 4
 ¼ 0,
ðRC Þ2 LC
340 Appendix. Solutions to Exercise Problems

L
R2 ¼ :
4C
The condition for an overdamped circuit is when the two solutions are two
different negative real numbers. First, we require

1 4
 > 0,
ðRC Þ 2 LC

L
R2 < :
4C
The condition for an underdamped circuit is when the two solutions are a pair of
complex conjugate numbers.

1 4
 < 0,
ðRC Þ2 LC
L
R2 > :
4C
When the circuit is overdamped or critically damped, the stability requirement is
that the solutions be negative, that is,
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 1 4
 þ  < 0,
RC ðRC Þ2 LC
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 4 1
 < ,
ðRC Þ2 LC RC

1 4 1
 < ,
ðRC Þ2 LC ðRC Þ2
4
0< ,
LC
which is always valid.
Our conditions are the same as those for the parallel RLC circuits.

Chapter 26. Filters

Problem 26.1 Without doing any mathematical derivation, determine whether the
Sallen-Key filter shown in Fig. P26.1 a lowpass filter or a highpass filter.
Appendix. Solutions to Exercise Problems 341

Fig. P26.1

Solution
Let us consider two extreme cases.

When the frequency is very low, the capacitors are open circuits. Figure P26.1
becomes Fig. S26.1a.

Fig. S26.1a

This circuit is a voltage follower (also known as a buffer) with

vout ¼ vin :

When the frequency is very high, the capacitors are short circuits. Figure P26.1
becomes Fig. S26.1b.
342 Appendix. Solutions to Exercise Problems

Fig. S26.1b

In Fig. S26.1b, the non-inverting input of the op-amp is shorted to the ground. As
a result, the inverting input and the output also have the voltage of 0.
Therefore, this is a lowpass filter.

Problem 26.2 Without doing any mathematical derivation, determine whether the
Sallen-Key filter shown in Fig. P26.2 a lowpass filter or a highpass filter.

Fig. P26.2

Solution
Let us consider two extreme cases.
Appendix. Solutions to Exercise Problems 343

When the frequency is very low, the capacitors are open circuits. Figure P26.2
becomes Fig. S26.2a.

Fig. S26.2a

In this circuit, the input is not fed into the circuit. The op-amp’s inverting and
non-inverting inputs are 0. Therefore, the output is 0.
When the frequency is very high, the capacitors are short circuits. Figure P26.2
becomes Fig. S26.2b.

Fig. S26.2b

Figure S26.2b is a voltage follower,


344 Appendix. Solutions to Exercise Problems

vout ¼ vin :

Therefore, this is a highpass filter.

Problem 26.3 Without doing any mathematical derivation, determine whether the
Sallen-Key filter shown in Fig.P26.2 a lowpass filter or a highpass filter or none
of them.

Fig. P26.3

Solution
At very low frequency, the capacitors are open circuits. Figure P26.3 becomes
Fig. S26.3a.

Fig. S26.3a
Appendix. Solutions to Exercise Problems 345

In this circuit, the op-amp’s inverting and non-inverting inputs are 0. Therefore,
the output is 0.
At very high frequency, the capacitors are short circuits. Figure S26.3a becomes
Fig. S26.3b.

Fig. S26.3b

Once again, in Fig. S26.3b, the op-amp’s inverting and non-inverting inputs are
0. Therefore, the output is 0.
This filter has 0 output at very low frequency and very high frequency. It is not a
lowpass filter and is not a highpass filter, either. We observe that there are two
capacitors in the circuit. This circuit is most likely a second-order bandpass filter.

Chapter 27. Wrapping Up

Problem 27.1 Are KVL equations and mesh equations the same? Are KCL
equations and node equations the same.

Solution
Mesh equations are KVL equations and are the special applications of the KVL
principle. A circuit may have many elements. If we use the element’s currents and
voltages as variables, we need a lot of equations to solve them. In principle, if we
have 20 variables, we need 20 equations.

A circuit may only have few meshes. Thus, we only need to solve a system of a small
number of mesh equations. Once the mesh currents are obtained, the element’s
currents and voltages can be readily evaluated using the mesh currents.
346 Appendix. Solutions to Exercise Problems

Similarly, node equations are the special applications of the KCL principle. A
circuit may have few essential nodes. An essential node is a node joining three or
more elements. Thus, we only need to solve a system of a small number of node
equations. Once the node voltages are obtained, the element’s currents and voltages
can be readily evaluated using the node voltages.
The biggest motivation to use node equations or mesh equations is to reduce the
number of equations to its minimum.

Problem 27.2 Do you have a preference regarding to nodes equations or mesh


equations?

Solution
In principle, if the number of essentials nodes is greater than the number of meshes,
use the mesh equation; otherwise, use the node equations. You commonly have
enough equations to solve for the circuit. You do not need to use both mesh
equations and node equations to start with circuit analysis.

There are many other strategies. One strategy is only set up equations for the
inductor currents and capacitor voltages. Once the inductor currents and capacitor
voltages are obtained, other currents and voltages can be easily solved.

Problem 27.3 The main purpose of the Laplace transform and the Fourier transform
is to avoid solving differential equations in the time domain. Why do we need both
the Laplace transform and the Fourier transform? Can we just learn one of them?

Solution
The applications of the Laplace transform and Fourier transform have many
overlaps. In many cases, you can use either the Laplace transform or the Fourier
transform. By using Laplace transform and Fourier transform, you can treat
capacitors and inductors in the same way as you treat resistors.

The Laplace transform is easier to work with if the circuit involves initial
conditions and switch actions. The Laplace transform is easier to use for time
transient analysis.
The Laplace transform assumes that the signals are zero when t < 0. On the other
hand, the Fourier transform does not have this restriction. The Fourier transform
assumes zero initial conditions. If we have to use the Fourier transform to deal with a
circuit with initial conditions, we need to artificially introduce some step sources into
the capacitors and inductors.

Problem 27.4 Which is more powerful, the Fourier transform method or the phasor
method?
Appendix. Solutions to Exercise Problems 347

Solution
The Fourier transform method is more powerful and has wider applications. The
phasor method can only handle one frequency. The Fourier transform method can
solve all the problems that the phasor method can solve. On the other hand, the
phasor method is easier to use in applications where the frequency is always fixed,
for example, in the power system.

Problem 27.5 Name one most important concept in electric circuits.

Solution
In our opinion, it is Ohm’s law.
Bibliography (Some Textbooks Used
in Colleges)

Basic Engineering Circuit Analysis, 11th Ed., J. David Irwin and R. Mark Nelms,
John Wiley and Sons. ISBN: 978-1-118-53929-3, 2015.
Circuit Analysis and Design, Fawwaz T. Ulaby, Michel M. Maharbiz and Cynthia
M. Furse, Michigan Publishing, 2018
Electric Circuits, 11th Edition, James W. Nilsson and Susan Reidel, Pearson, ISBN-
139780134746968, 2019
Engineering Circuit Analysis, 9th Edition, William Hayt, Jack Kemmerly, and
Steven Durbin, McGraw Hill, ISBN13: 9780073545516, 2019.
Essentials of Electrical and Computer Engineering, David V. Kerns, Jr. and J. David
Irwin, Pearson Prentice Hall, Upper saddle River, ISBN-13: 978-0139239700,
2004
Fundamentals of Electric Circuits, 6th edition, Charles K. Alexander and Matthew
N. O. Sadiku, Mc Graw Hill, ISBN-13: 978-0078028229, 2017
Fundamentals of Electronic Circuit design, David J. Comer and Donald T. Comer,
Wiley, ISBN-13: 978-0471410164, 2002
Introduction to Electric Circuits, James A. Svoboda and Richard C. Dorf, John
Wiley & Sons Inc., NY, 9th Edition, 2013. ISBN 1118477502,
The Analysis and Design of Linear Circuits, 7th edition, Roland E. Thomas, Albert
J. Rosa and Gregory J. Toussaint, John Wiley & Sons, Inc.
The Analysis and Design of Linear Circuits (electronic version OK), Roland
E. Thomas and Albert J. Rosa and Gregory J. Toussaint, Wiley, ISBN 978-1-
118-06558-7, 2012.

# The Editor(s) (if applicable) and The Author(s), under exclusive license to 349
Springer Nature Switzerland AG 2021
G. L. Zeng, M. Zeng, Electric Circuits, https://doi.org/10.1007/978-3-030-60515-5
Index

A Dot convention, 150


Active filter, 203 Dot product, 160
Alternating current (AC), 130
Ammeter, 11
Amplitude, 134 E
Eigenfunction, 129
Energy, 105, 114
B Equivalent circuit, 81
Bandpass filter, 205 Essential node, 49
Breadboard, 9 Euler’s formula, 132
Even function, 163

C
Capacitance, 111 F
Capacitor, 111 Filter, 203
Characteristic equation, 189 Final value, 126
Comparator circuit, 97 Final voltage value, 111
Conductance, 40 First-order circuits, 125
Controlled source, 74 Fourier coefficients, 157
Coupling, 143 Fourier series, 157
Cover-up method, 176 Fourier transform, 181
Critically damped, 187 Function generator, 137
Current, 2 Fundamental frequency, 157
Current divider, 44
Current-voltage characteristic curves (I–V
curves), 4 G
Galvanometer, 44
Ground, 1
D
dB, 205
DC current source, 4 H
DC offset, 163 Half-wave symmetry, 163
DC power supply, 9 Harmonic frequencies, 157
DC voltage source, 4 Highpass filter, 202
Decibel, 205
Delta function, 173
Dependent source, 74 I
Differential equations, 119–122 Ideal transformer, 151
Dirac delta function, 173 Impedance, 94, 130

# The Editor(s) (if applicable) and The Author(s), under exclusive license to 351
Springer Nature Switzerland AG 2021
G. L. Zeng, M. Zeng, Electric Circuits, https://doi.org/10.1007/978-3-030-60515-5
352 Index

Impedance matching, 94, 151 Parseval's theorem, 166


Independent source, 73 Partial fraction decomposition, 175
Inductance, 105 Partial fraction expansion, 175
Inductor, 105 Passive filter, 203
Initial value, 126 Passive sign convention, 3
Initial voltage value, 111 Period, 157
Inner product, 160 Periodic function, 157
Internal resistance, 4 Phase, 134
Inverse phasor transform, 135, 159 Phasor, 129
Inverting amplifier, 98, 204 Phasor transform, 135
Power, 93, 105, 114
Power rails, 9
K Probe, 144
Kirchhoff’s current law (KCL), 31
Kirchhoff’s voltage law (KVL), 23
R
Reflected impedance, 152
L Resistance, 3, 38
Laplace transform, 171 Resistor, 4
Laplace transform pairs, 172 Resonance radian frequency, 205
Loop, 24
Lowpass filter, 201
S
Second-order circuit, 187
M Series, 37, 108, 113
Maximum power transfer, 93 Short circuit, 4
Mesh, 59 Signum function, 181
Mesh-current method, 59 Step response, 187
Multimeter, 10 Supermesh, 62
Multisim, 67, 71, 72, 260–271 Supernode, 32
Mutual inductance, 149

T
N Thévenin equivalent circuit, 81
Negative feedback, 97 Time constant, 106, 111, 126
Node, 32 Transfer function, 210
Node-voltage method, 49 Trigger, 144
Non-inverting amplifier, 99
Norton equivalent circuit, 84
U
Under damped, 187
O Unit impulse function, 173
Odd function, 163 Unit step function, 173
Ohm’s law, 17, 18
Op amp, 97
Open circuit, 4 V
Operational amplifier, 97 Voltage, 1
Orthogonal, 160 Voltage divider, 43
Orthogonality, 160 Voltage follower, 99
Oscilloscope, 137 Voltmeter, 11
Over damped, 187

W
P Wheatstone bridge, 44
Parallel, 37, 108, 113

You might also like