You are on page 1of 30

BASIC ELECTRONICS ENGINEERING

Jay-Cee R. Billones
VISION
Laguna University shall be a socially responsive educational
institution of choice providing holistically developed
individuals in the Asia-Pacific Region.

MISSION
Laguna University is committed to produce academically
prepared and technically skilled individuals who are socially
and morally upright citizens.

Department of Mechanical Engineering


MISSION
The Department of Mechanical Engineering of Laguna
University is committed to produce academically prepared
and technically skilled mechanical engineers who are socially
and morally upright citizens.
VISION
The Department of Mechanical Engineering of Laguna
University is envisioned to be the provincial college of choice
producing well-equipped mechanical engineers who
specializes on energy management.
Table of Contents

Module 1: An Introduction To Circuit Engineering 1


Introduction 1
Learning Objectives 1
Lesson 1. Circuit and its parts 2
Assessment Task 7
Summary 7
References 7

Module 2: Basic Concepts And Resistor Circuits 8


Introduction 8
Learning Objectives 8
Lesson 1. Basic Concept 9
Lesson 2. Resistors 12
Lesson 3. AC Signals 17
Assessment Task 20
Summary 21
Reference 21

Module 3: AC Circuits 22
Introduction 22
Learning Objectives 22
Lesson 1. Capacitors 23
Lesson 2. Inductors 25
Assessment Task 26
Summary 26
References 26
Course Code: Eng’g 202

Course Description: This course discusses the construction, operation and


characteristics of basic electronics devices such as PN junction diode, light emitting
diode, Zener diode, Bipolar Junction Transistor and Field Effect Transistor. Diode
circuit applications such as clipper, clamper and switching diode circuits will be a
part of the lecture. Operation of a DC regulated power supply as well as analysis of
BJT and FET amplifier circuit will be tackled. This course also discusses the
operation and characteristics of operational amplifiers

Course Intended Learning Outcomes (CILO):


After completing this course, the student must be able to:
1. Understand the basic operation, construction and characteristics of
different electronic devices such as PN, junction diode, light emitting
diode, Zener diode, Bipolar junction Transistor, Field Effect Transistor and
Operational Amplifier as well as their application
2. Understand the operation of a DC regulated power supply.
3. Analyze BJT and FET amplifier circuits.
4. Analyze Operational amplifier circuits.
Course Requirements:
Assessment Tasks - 60%
Major Exams - 40%
_________
Periodic Grade 100%

Computation of Grades:

PRELIM GRADE = 60% (Activity 1-4) + 40% (Prelim exam)

MIDTERM GRADE = 30%(Prelim Grade) + 70 %[60% (Activity 5-7) + 40% (Midterm exam)]

FINAL GRADE = 30%(Midterm Grade) + 70 %[60% (Activity 8-10) + 40% (Final exam)]
MODULE 1
AN INTRODUCTION TO CIRCUIT
ENGINEERING

Introduction

According to Solis (2015), in 1882, there was a circuit war; it was between the notable
electrical engineers and scientists, Thomas Edison (inventor of the DC system) and Nikola Tesla
(inventor of the AC system).

While Thomas Edison stated that an efficient way of distributing power was via a DC
system, Nikola Tesla argued that although DC systems are efficient, an alternating current is the
more practical option. It started as a simple clash of ideas, but it eventually led to a major rift.
Neither professional conceded; both of them insisted that their own systems were “better” (Solis,
2015).

Learning Outcomes

At the end of this module, students should be able to:

1. Know about the history of circuit engineering


2. Familiarized with the basics of electronics

1
Lesson 1. Circuit and its parts

What is a Circuit?
Both an electric circuit and an electronic circuit refer to a complete pathway for electric
current, which starts and ends at a single point; it is a passage that allows the electricity to enter
at one place, then, let it pass through a series of stops, and finally, leave it to exit at the same
place. The list of basic examples of a circuit includes a light switch (off and on) and battery-
operated lamps (Solis, 2015).

Figure 1.1. A circuit that follows a fundamental design

A circuit can function well - granted that its design is well-conceptualized. As much as
possible, it is recommended that arriving at a simplistic product should be the goal; the simple
and straightforward a design is, the better. With a fundamental concept, even if other (beginner-
level) circuit engineers who will subject it to inspection will not have a difficult time in
understanding its flow. Although there may be complex systems, the agenda is not intended to
complicate the explanations (Solis, 2015).

2 classifications of a circuit (Solis, 2015):

1. Linear or non-linear – a circuit that is based on either linear or non-linear networks; it is


composed of independent and/or dependent sources and passive elements

2
2. Active or passive – a circuit that is based on either the absence (passive circuit) or the
presence (active circuit) of a source; a source can be a power source or voltage source

A Circuit & Its Types

Not all circuits are alike. In fact, one of the most common misconceptions involves an
electric circuit and an electronic circuit; both are said to be one and the same, but they are not.
While the former can carry average to high voltage, the latter has the tendency to have low voltage
load (Solis, 2015).
Circuit types (Solis, 2015):
Closed circuit – it is a circuit that is fully functional
 Open circuit – it is a circuit that can no longer function due to a damaged or missing component,
or a loose connection
 Short circuit – it is a circuit that comes without a load
 Parallel circuit – it is a circuit that connects to other circuits; it is like the main power source or
the primary circuit in a series of circuits
 Series circuit – it is a circuit that connects to other circuits; the same amount of electricity is
distributed to each of its component circuits; the main power source or the primary circuit is
unclear

Conductors, Insulators & Semi-Conductors

Conductors, insulators, and semi-conductors give light to the fact that a circuit’s electrical
properties are dependent on the circuit type, as well as on their conduction bands (i.e. their
allowed electric power). For instance, if a particular power source chooses to distribute a 9-volt
electric power to a closed circuit, its electrical properties can be evaluated by using 2 details: (1)
its characteristic as a closed circuit and (2) 9-volt electric power (Solis, 2015).
Moreover, conductors, insulators, and semi-conductors are integral concepts to the
conductivity of an object. While conductors and semi-conductors are grouped to describe charged
carriers, insulators are still considered as relative despite not containing any free charge (Solis,
2015).

3
Figure 1.2. Valance bond of Insulators, semi-conductors, conductors

Conduction bands (Solis, 2015):


 Conductor – it is a conduction band that is referred to as the almost full band
 Insulator – it is a conduction band that is referred to as an empty band
 Semi-conductor –it is a conduction band that is referred to as an almost empty band

Breaking Down the Components of a Circuit

A circuit can be either simple or complex, and be both simple and complex. If the circuit in
subject is a series circuit, with a group of 10 different circuits that are connected to it or if the said
circuit is just a basic closed circuit with 5 different stops, it can be rather confusing to trace.
However, if you dissect any circuit, you’ll discover 3 constant, integral components (Solis, 2015).

Integral components (Solis, 2015):


1. Load – it is the representation of the power consumption, as well as the work that is
accomplished within a system; without it, there’s barely a point in having a circuit
2. Power source – it is where the electricity comes from.
3. Pathway – it is the framework of a circuit; from the power source, it follows the load through
each of the network, and finally returns to and exits the power source; it is also referred to as
the conductive pathway

4
The Roles of Current, Resistance & Voltage
Current, resistance, and voltage are the 3 representations of the important components of
a circuit’s system. They can explain how electricity enters, then, moves from 1 point to another,
and finally exits. Whether the path of electricity is rather simple, these representations remain
constant. Apart from describing the electric flow, they can serve as indications of faults (in
instances when a circuit fails to work) (Solis, 2015).
3 representations (Solis, 2015):
 Current – it is the representation of the electric flow; particularly, its focus is on the flow of
electrons
 Resistance – it is the representation of the nature of an electric flow as it moves around the
circuit
 Voltage – it is the representation of the electric force or pressure; in general, the supply comes
from an electric outlet or a battery

AC/DC Systems (Solis, 2015)

AC and DC systems (or alternating current and direct current systems) are often
associated to each other. When the AC system is mentioned, so is the DC system. Conversely,
when it’s the DC system’s turn to be in the spotlight, it won’t be long until the AC system is
mentioned. This is because these systems are opposite of one another; to get a better
understanding of one of them, it’s recommended to be familiar with the other, as well.
Moreover, AC and DC systems are types of a circuit’s current flow. In an AC system, the
current flow changes its direction occasionally. Meanwhile, in a DC system, the current flow
follows a single direction.
It can be deduced, therefore, that an AC system grants a circuit freedom to let the current
flow in several directions. While this can be an advantage, this doesn’t permit the continuous flow
that a DC system can entitle.
So, should you use an AC system or a DC system? The decision as to which current
system is dependent on the more practical design to follow; take into account the aim of having
your own circuit. If you prefer something grand and you intend to power something large, the AC
system can step in. On the other hand, if you’re good with a basic setup, you can use the DC
system’s concept as basis.

5
What Is a Transformer?

A transformer is a device that serves as a portal for energy transfer within the points in a
circuit or from circuit 1 to circuit 2. In most cases, it is used for increasing and decreasing the
voltages in a system (Solis, 2015).
When the first transformer was built in the mid 1880’s, circuit engineers discovered that a
transformer significantly improves the electric flow in a circuit, and consequently, results to a more
powerful circuit. The discovery made way for various transformer designs, as well as various
transformer sizes (Solis, 2015).
A primary principle of a transformer is its need for extremely high magnetic permeability.
It follows that a circuit that is capable of attracting power is, of course, more inclined to have
electric current transferred to it; and, conversely, a circuit with low magnetic permeability is less
likely to extract power from another circuit (Solis, 2015).

Figure 1.3. Magnetic permeability is defined as the ability of a circuit to hold and support an
internal magnetic field

6
Assessment Task 1

1. Create an schematic diagram of the following circuit (open, short, parallel and series).
Include also the pathway, load and the power source (AC/DC).

2. Give at least 5 each material used as insulators, conductors and semi-conductors.

Summary

Current – it is the representation of the electric flow; particularly, its focus is on the flow of electrons
Resistance – it is the representation of the nature of an electric flow as it moves around the circuit
Voltage – it is the representation of the electric force or pressure; in general, the supply comes
from an electric outlet or a battery
Transformer is a device that serves as a portal for energy transfer within the points in a circuit or
from circuit 1 to circuit 2

Reference

7
Solis Tech. (2015). Circuit Engineering “The Beginner’s Guide to Electronic Circuits, Semi-
Conductors, Circuit Boards, and Basic Electronics”

MODULE 2
BASIC CONCEPTS AND RESISTOR CIRCUITS

Introduction

We start our study of electronics with definitions and the basic laws that apply to all circuits.
This is followed by an introduction to our first circuit element – the resistor (Eggleston, 2011).

In electronics, we are interested in keeping track of two basic quantities: the currents and
voltages in a circuit. If you can make these quantities behave like you want, you have succeeded
(Eggleston, 2011).

Learning Outcomes

At the end of this module, students should be able to:

1. Understand the importance of resistor in a circuit.


2. Know how to solve current, voltage and resistance.

8
Lesson 1. Basic Concept
Current measures the flow of charge past a point in the circuit. The units of current are
thus coulombs per second or amperes, abbreviated as A. In this text we will use the symbol I or i
for current (Eggleston, 2011).

As charges move in circuits, they undergo collisions with atoms and lose some of their
energy. It thus takes some work to move charges around a circuit. The work per unit charge
required to move some charge between two points is called the voltage between those points. (In
physics, this work per unit charge is equivalent to the difference in electrostatic potential between
the two points, so the term potential difference is sometimes used for voltage (Eggleston, 2011).

The simplest circuit will involve one voltage source and one sink, with connecting wires
as shown in Figure. 2.1. By convention, we denote the two sides of the voltage source as + and
−. A positive charge moving from the − side to the + side of the source gains energy. Thus we
say that the voltage across the source is positive. When the circuit is complete, current flows out
of the + side of the source, as shown. The voltage across the component is negative when we

Figure 2.1. A simple generic circuit.

cross it in the direction of the current. We say there is a voltage drop across the component. Note
that while we can speak of the current at any point in the circuit, the voltage is always between
two points. It makes no sense to speak of the voltage at a point (remember, the voltage is a
potential difference). We can now write down some general rules about voltage and current
(Eggleston, 2011).

9
1. The sum of the currents into a node (i.e. any point on the circuit) equals the sum of the
currents flowing out of the node. This is Kirchoff’s Current Law (KCL) and expresses
conservation of charge. For example, in Figure 2.2, I1 = I2 + I3. If we use the sign convention
that currents into a node are positive and currents out of a node are negative, then we can
express this law in the compact form

Figure 2.2. Example of Kirchoff’s Current Law

That is equal to:


𝑛𝑜𝑑𝑒

∑ 𝐼𝑘 = 0 eq′n 2.1
𝑘
where the sum is over all currents into or out of the node (Eggleston, 2011).
2. The sum of the voltages around any closed circuit is zero. This is Kirchoff’s Voltage Law
(KVL) and expresses conservation of energy. In equation form,
𝑙𝑜𝑜𝑝

∑ 𝑉𝑘 = 0 𝑒𝑞 ′ 𝑛 2.2
𝑘
Here we must use the convention that the voltage across a source is positive when we move
across the source in the direction of the current and the voltage across a sink is negative
when we move across the component in the direction of the current. If we traverse a source
or sink in the direction opposite to the direction of the current, the signs are reversed. Figure
2.3 gives an example. Here we introduce the circuit symbol for an ideal battery, labeled with
voltage V1. The top of this symbol represents the positive side of the battery (Eggleston,
2011).

10
Figure 2.3. Example of Kirchoff’s Voltage Law

The current (not shown) flows up out of the battery, through the component labeled V2 and
down through the components labeled V3 and V4. Looping around the left side of the circuit
in the direction shown gives V1 − V2 − V3 = 0 or V1 = V2 + V3. Here we take V2 and V3 to be
positive numbers and include the sign explicitly. Going around the right portion of the circuit
as shown gives −V3 + V4 = 0 or V3 = V4. This last equality expresses the important result that
components connected in parallel have the same voltage across them (Eggleston, 2011).
3. The power P provided or consumed by a circuit device is given by
P=IV eq’n 2.3
where V is the voltage across the device and I is the current through the device. This follows
from the definitions:
𝑤𝑜𝑟𝑘 𝑐ℎ𝑎𝑟𝑔𝑒 𝑤𝑜𝑟𝑘
𝑉𝐼 = ( )( ) = = 𝑝𝑜𝑤𝑒𝑟 𝑒𝑞 ′ 𝑛2.4
𝑐ℎ𝑎𝑟𝑔𝑒 𝑡𝑖𝑚𝑒 𝑡𝑖𝑚𝑒
The units of power are thus joules per second or watts, abbreviated W. This law is of
considerable practical importance since a key part of designing a circuit is to employ
components with the proper power rating. A component with an insufficient power rating will
quickly overheat and fail when the circuit is operated.
Some common prefixes and their meanings are shown in Table 2.1. As an example, recall
that the unit volts is abbreviated as V, and amperes or amps is abbreviated as A. Thus 10 6
volts = 1 MV and 10−3 amps = 1 mA. Notice that case matters: 1 MA ≠ 1 mA (Eggleston,
2011).

Figure 2.4. I-V curve for a resistor


Table 2.1. Common prefixes use in electronics

11
Lesson 2. Resistors
A common way to represent the behavior of a circuit device is the I–V characteristic. This is
a plot of the current I through the device as a function of applied voltage V across the device. Our
first device, the resistor, has the simple linear I–V characteristic shown in Fig. 1.4. This linear
relationship is expressed by Ohm’s Law: (Eggleston, 2011).
V=IR eq’n 2.3
The constant of proportionality, R, is called the resistance of the device and is equal to one
over the slope of the I–V characteristic. The units of resistance are ohms, abbreviated as Ω. Any
device with a linear I–V characteristic is called a resistor (Eggleston, 2011).
The resistance of the device depends only on its physical properties – its size and
composition. More specifically: (Eggleston, 2011).

eq’n 2.4
where ρ is the resistivity, L is the length, and A is the cross-sectional area of the material.
The resistivity of some representative materials is given in Table 2.2 (Eggleston, 2011).

Table 2.2. Electronic Material Resistivity

12
The interconnecting wires or circuit board paths are typically made of copper or some other
low resistivity material, so for most cases their resistance can be ignored. If we want resistance
in a circuit we will use a discrete device made of some high resistivity material (e.g., carbon).
Such resistors are widely used and can be obtained in a variety of values and power ratings
(Eggleston, 2011).

The low power rating resistors typically used in circuits are marked with color coded bands
that give the resistance and the tolerance (i.e., the uncertainty in the resistance value) as shown
schematically in Figure 2.5 (Eggleston, 2011).

Figure 2.5 Value and tolerance bands on a resistor

As shown in the figure, the bands are usually grouped toward one end of the resistor. The
band closest to the end is read as the first digit of the value. The next band is the second digit,
the next band is the multiplier, and the last band is the tolerance value. The values associated
with the various colors are shown in Table 2.3. For example, a resistor code having colors red,
violet, orange, and gold corresponds to a value of 27 × 103 Ω ± 5% (Eggleston, 2011).

Resistors also come in variable forms. If the variable device has two leads, it is called a
rheostat. The more common and versatile type with three leads is called a potentiometer or a
“pot.” Schematic symbols for resistors are shown in Figure 2.6 (Eggleston, 2011).

Table 2.3 Standard color scheme for Resistor

13
Figure 2.6. Schematic symbols for a fixed resistor and two types of variable resistors

As noted in Eq’n. 2.3, the power consumed by a device is given by P = VI, but for resistors
we also have the relation V = IR. Combining these we obtain two power relations specific to
resistors: P = I 2R ; P = V 2/ R eq’n 2.5 (Eggleston, 2011).

2.1 Resistors in series (Eggleston, 2011)


Components connected in series are connected in a head-to-tail fashion, thus forming a line
or series of components. When forming equivalent circuits, any number of resistors in series may
be replaced by a single equivalent resistor given by:

Figure 2.7 Equivalent circuits for a resistor in series


Where

eq’n 2.6

By KCL, the current in each resistor is the same. Applying KVL around the circuit loop and
Ohm’s Law for the drop across the resistors, we obtain

V = IR1 + IR2 + IR3 ; V = I(R1 + R2 + R3); V = IReq Eq’n 2.7

Req = R1 + R2 + R3

14
2.2 Resistor in Parallel (Eggleston, 2011)

Components connected in parallel are connected in a head-to-head and tail-to-tail fashion. The
components are often drawn in parallel lines, hence the name. When forming equivalent circuits,
any number of resistors in parallel may be replaced by a single equivalent resistor given by:

eq’n 2.8

where the sum is over all the resistors in parallel. To see this, consider the circuit shown in Figure
2.8. Again, we would like to replace the circuit on the left by the equivalent circuit on the right
(Eggleston, 2011).

Figure 2.8. Equivalent circuit for resistors in parallel

First, note that KCL requires I = I1 + I2 + I3 eq’n 2.9.


Since the resistors are connected in parallel, the voltage across each one is the same, and, by
KVL is equal to the battery voltage: V = I1R1, V = I2R2, V = I3R3. Solving these for the three currents
and substituting in Eq’n. 2.9 gives

Eq’n 2.10
Where

For special case is worth memorizing

eq’n 2.11

Example problem 2.1. Resistor and Power rating (Eggleston, 2011)

15
For the circuit shown in Figure 2.9, how much current flows through the 20kΩ resistor? What must
its power rating be?

Figure 2.9. Example resistor circuit


Solution:

As we will see, there is more than one way to solve this problem. Here we use a method
that relies on basic electronics reasoning and our resistor equivalent circuit laws. We want the
current through the 20kΩ resistor. If we knew the voltage across this resistor (call this voltage
V20k), we could then get the current from Ohm’s Law. In order to get the voltage across the 20kΩ
resistor, we need the voltage across the 10kΩ resistor since, by KVL, V20k = 130 − V10k. In order
to get the voltage across the 10kΩ resistor, we need to know the current through it, which is the
same as the current supplied by the battery. Thus, if we can get the current supplied by the battery
we can solve the problem. To get the battery current, we combine all our resistors into one
equivalent resistor. The implementation of this strategy goes as follows.

1. Combine the two 5 k series resistors into a 10 k resistor.


2. This 10 k resistor is then in parallel with the 20 k resistor. Combining thesewe get (using
Eq’n 2.11) Req = 6.67kΩ
3. This 6.67kΩ resistor is then in series with a 10kΩ resistor, giving a total equivalent circuit
resistance Req = 16.67kΩ.
𝑉𝑜
4. The current supplied by the battery is then 𝐼 = = 7.8𝑚𝐴
𝑅𝑒𝑞

5. KVL then gives 130V − (7.8 mA)(10kΩ) − V20k = 0. Solving this gives V20k = 52V.
6. Ohm’s Law then gives I20k = 52V/20kΩ = 2.6 mA, which is the solution to the first part of
our problem. As a check, it is comforting to note that this current is less than the total
battery current, as it must be. The remainder goes through the two 5kΩ resistors.
7. The power consumed by the 20kΩ resistor is P = I2R = (2.6 × 10−3 A)2(2 ×104 Ω) = 0.135
W. This is too much for a 1/8 W resistor, so we must use at least a 1/4 W resistor.

16
Lesson 3. AC Signals (Eggleston, 2011)

So far our examples have used constant voltage sources such as batteries. Constant
voltages and currents are described as DC quantities in electronics. On the other hand, voltages
and currents that vary in time are called AC quantities. For future reference, we list here some of
the most common AC signals (Eggleston, 2011).

Figure 2.10. A sine wave

1. Sinusoidal signals. This is probably the most fundamental signal in electronics since, as we
will see later, any signal can be constructed from sinusoidal signals (Eggleston, 2011). A
typical sinusoidal voltage is shown in Figure 2.10. Sinusoidal voltages can be written
V = A sin(2πft + Φ) = A sin(ωt + Φ) eq’n 2.12

Where:

A – Amplitude; f – frequency in cycle per seconds or hertz (Hz); Φ – phase; ω – angular frequency
in radian per second; t – time in repetition or T in signal that is related to frequency of the signal
by T = 1/f.

There are several ways to specify the amplitude of a sinusoidal signal that are in common use.
These include the following.
(a) The peak amplitude A or Ap.

(b) The peak-to-peak amplitude App = 2A.

(c) The rms amplitude Arms = A /√2. This is useful for power calculations involving sinusoidal
waves. For example, suppose we want the power dissipated in a resistor given the sinusoidally
varying voltage across it. We cannot simply use Eq’n 2.5 since our voltage is varying in time (what

17
V would we use?). Instead, we calculate the time average of the power over one period:

eq’n 2.13

This last form shows that we can use Eq. (1.8) to calculate the power as long as we use the rms
amplitude of the sinusoidal signal in the formula.

(d) Decibels (abbreviated dB) are used to compare the amplitude of two signals,

say A1 and A2:

eq’n 2.14

where this last expression uses the power level of the two signals. So, for example, if A2 = 2A1,
then we get 20 log 2 ≈ 6, so we say A 2 is 6 dB higher than A1. Various related schemes specify
the decibel level relative to a fixed standard. So dBV is the dB relative to a 1 Vrms signal and dBm
is the dB relative to a 0.78 Vrms signal. For the curious, this latter voltage standard is 1 mW into a
600Ω resistor.

Some other typical waveforms of electronics are shown in Figures 2.11 through 2.16.

2. Square wave. Specified by an amplitude and a frequency (or period) (Eggleston, 2011).

Figure 2.11. The square wave

3. Sawtooth wave. Specified by an amplitude and a frequency (or period) (Eggleston, 2011).

18
Figure 2.12. The sawtooth wave

4. Triangle wave. Specified by an amplitude and a frequency (or period) (Eggleston, 2011).

Figure 2.13. The triangle wave

5. Ramp. Specified by an amplitude and a ramp time (Eggleston, 2011).

Figure 2.14. A ramp signal

6. Pulse train. Specified by an amplitude, a pulse width τ , and a repetition time trep. The duty
cycle of a pulse train is defined as τ/trep (Eggleston, 2011).

Figure 2.15. A pulse train

19
7. Noise. These are random signals of thermal origin or simply unwanted signals coupled into
the circuit. Noise is usually described by its frequency content, but that is a more advanced
topic (Eggleston, 2011).

Figure 2.16. Noise

Assessment Task 2

1. Compute the current through R2 and R3 of figure below

2. Compute the current through R1 and R2 of Figure below

20
Summary

According to Eggleston (2011), in electronics, we are interested in keeping track of two


basic quantities: the currents and voltages in a circuit. Current measures the flow of charge past
a point in the circuit. The units of current are thus coulombs per second or amperes, abbreviated
as A. In this text we will use the symbol I or i for current. The work per unit charge required to
move some charge between two points is called the voltage between those points. The units of
voltage are thus joules per coulomb or volts, abbreviated V.

Reference

Eggleston, Dennis L. (2011). Basic Electronics for Scientists and Engineers. Cambridge
University Press, The Edinburgh Building, Cambridge CB2 8RU, UK

21
MODULE 3
AC CIRCUITS

Introduction

According to Eggleston (2011), currents and voltages that vary in time are called AC
quantities. When analyzing circuits where the current and voltage change in time, the treatment
of resistors is unchanged: they still obey Ohm’s Law. In this chapter we introduce two other basic
circuit components, the capacitor and the inductor. The treatment of these components depends
on the details of how things are changing in time, and this will require the development of some
new analysis techniques.

Learning Outcomes

At the end of this module, students should be able to:

1. Know about the AC circuits


2. Learn how to compute capacitors and inductors

22
Lesson 1. Capacitors (Eggleston, 2011)

Another basic circuit component is the capacitor. A capacitor is formed by any pair of
conductors, but the usual form is two parallel plates. For this case, the capacitance C is given by
𝐴
𝐶 = 𝜖 ; (F) eq’n 3.1
𝑑

Where

A - is the area of a plate, d is the distance between plates

∈ - is the dielectric constant of the material between the plates. Note that, like the resistance, the
capacitance depends only on the physical characteristics of the device. The unit of capacitance
is coulombs per volt or farads, abbreviated F. Typical capacitor values are in a range such that
μF or pF are convenient units. When purchasing a capacitor, you must specify its voltage rating
in addition to its capacitance value. This rating tells you the maximum voltage you can apply
across the capacitor before there is electrical breakdown through the dielectric material
(Eggleston, 2011).

So what does a capacitor do? One answer is that it is a charge storage device. When a
voltage V is applied to a capacitor, a charge of magnitude Q will be stored on each plate. Q is
given by
Q = CV eq’n 3.2

Figure 3.1. Equivalent circuit capacitors in series

In electronics, we are usually concerned with currents (the flow of charge) rather than charge. If
we take the time derivative of Eq. (2.2) and note that, by definition, I =dQ/dt we get
𝑑𝑄
𝐼= 𝐶 eq’n 3.3
𝑑𝑡
Viewed from this perspective, C is the constant relating a time-varying voltage across the
capacitor to the AC current through the capacitor (Eggleston, 2011).

23
SERIES CAPACITORS (Eggleston, 2011)
Consider, for example, three capacitors in series as shown in Figure 3.1. We wish to
combine the capacitors to form the equivalent circuit on the right.
Let Q1 be the charge on capacitor C1 and so on. Applying KVL and Eq’n. 3.2 we obtain

eq’n 3.4
By charge conservation, the charge on each capacitor is the same, so Q1 = Q2 = Q3 ≡ Q and

eq’n 3.5
Comparing this with Eq’n 3.2, we see that the equivalent capacitance Ceq will be given by

eq’n 3.6
or, generalizing this to any number of capacitors in series,

eq’n 3.7
PARALLEL CAPACITORS (Eggleston, 2011)
Now consider three capacitors in parallel as shown in Figure 3.2. Again, let Q1 be the
charge on capacitor C1 and so on. Because the capacitors are connected in parallel, the voltage
across them must be the same

eq’n 3.8

Figure 3.2. Equivalent circuit for capacitors in parallel

If we add the three charges and apply Eq’n 3.8 to each term, we get
Q1 + Q2 + Q3 = V(C1 + C2 + C3) eq’n 3.9

24
If we are to form an equivalent capacitor the battery must supply the same amount of charge in
both cases. Thus Qeq = Q1 + Q2 + Q3 and, from Eq’n 3.9, Qeq =V(C1 + C2 + C3). Comparing this
with Eq’n 3.2, we see that
Ceq = C1 + C2 + C3 eq’n 3.10
or, generalizing this to any number of capacitors in parallel,

eq’n 3.11
Lesson 2. Inductors (Eggleston, 2011)

We learn in introductory physics that currents produce magnetic fields (Ampère’s Law)
and that time-varying magnetic fields can induce a voltage in a circuit (Faraday’s Law). Putting
these together meaning that time-varying currents in a circuit induce voltages (Eggleston, 2011).
This is expressed in equation form by

eq’n 3.12
where the constant L is called the self-inductance or simply the inductance. While any circuit loop
has inductance, we usually ignore this (like we ignore the small resistance of connecting wires)
and, if inductance is required in a circuit, add discrete inductors made of coils of wire (Eggleston,
2011). For a long coil (i.e., a solenoid), this inductance is given by

eq’n 3.13
where μ is the permeability of the material on which the coil is wound, N is the number of turns in
the coil, R is the radius of the coil, and l is the length of the coil. The unit of inductance is volts
times seconds per amp or henries (abbreviated H) (Eggleston, 2011).
The derivation of the equivalent circuit laws for inductors in series and parallel is similar to
that for resistors, and we leave the details to the reader. The result for inductors in series is

eq’n 3.14
and for inductors in parallel

eq’n 3.15

25
Assessment Task 3 (Eggleston, 2011)

1. Find the equivalent capacitance across the terminals of the circuit in figure
below

2. what value of C is needed so that V out = 70 V at 10.0s after the switch closes?
and taking V0 = 100 V, R = 1MΩ,

Summary
A capacitor is formed by any pair of conductors, but the usual form is two parallel plates. And
Inductors is currents produced magnetic fields (Ampère’s Law) and that time-varying magnetic
fields can induce a voltage in a circuit (Faraday’s Law) (Eggleston, 2011).

Reference

Eggleston, Dennis L. (2011). Basic Electronics for Scientists and Engineers. Cambridge
University Press, The Edinburgh Building, Cambridge CB2 8RU, UK

26

You might also like