[go: up one dir, main page]

0% found this document useful (0 votes)
2K views109 pages

Electrical Engineering Science 1 - 230419 - 164951

1. The document discusses the basics of electrical engineering science and the structure of atoms. 2. It covers topics like atoms, atomic structure, electrons, protons, neutrons, and how electrons are arranged in different orbits around the nucleus. 3. The course aims to help students understand concepts like electric current flow, simple DC circuits, different types of energy and relationships, electrostatics, electric charge, and capacitance.

Uploaded by

Akogun Elizabeth
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)
2K views109 pages

Electrical Engineering Science 1 - 230419 - 164951

1. The document discusses the basics of electrical engineering science and the structure of atoms. 2. It covers topics like atoms, atomic structure, electrons, protons, neutrons, and how electrons are arranged in different orbits around the nucleus. 3. The course aims to help students understand concepts like electric current flow, simple DC circuits, different types of energy and relationships, electrostatics, electric charge, and capacitance.

Uploaded by

Akogun Elizabeth
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/ 109

ELECTRICAL ENGINEERING SCIENCE 1

EEC 115

YEAR 1 FIRST SEMESTER


ND IN ELECTRICAL/ELECTRONIC ENGINEERING
TECHNOLOGY

LECTURE NOTE

BY

Engr. UWAGBOE J. O.

1
INTRODUCTION

Electrical Engineering Technology is an engineering technology field that implements and applies
the principles of electrical engineering. Electrical Technology deals with the design, application,
installation, manufacturing, operation and/or maintenance of electrical / electronic(s) systems. It
is a specialized discipline that has more for on application theory and applied design, electronics,
embedded systems, control systems, instrumentation, telecommunication and power system.
The course will cover the following topics;
 Atom
 The Structure and Composition of Atoms
 Conductors, Insulators and Semi-Conductors
 Current and Electron flow
 Electric Current, Potential Difference, Electromotive force (E.M.F), Resistance
 Electric Quantities
 D.C. Current
 Ohm’s Law
 Resistivity and Conductivity of a Conductor
 Series and Parallel circuits
 Kirchoff's laws.
 Superposition Principles
 Temperature Coefficient of resistance
 Types of energy
 Joule's law
 Energy Generations
 Electric charge
 Coulomb's law
 Define Electric Field Strength, Electric Flux Density, Permitivity, Relative Permitivity,
Field Intensity, Potential and Electric Flux.
 Capacitance

OUTCOMES:

2
On completion of the module, the students should be able:
1. Understand the concept of the electric current flow.
2. Understand simple D.C. circuits.
3. Know various types of energy and their inter-relationships.
4. Understand the concept of electrostatics, electric charge and capacitance of capacitor.

Assessment: The practical class will be awarded 40% of the total score. The continuous
assessments, test and quizzes will take 10% of the total score, while the remaining 50% will be for
the end of the semester examination score

3
MODULE 1

1.1 ATOMS
An atom is basically present in everything visible around us. Every living organism is composed of
atoms. Non-living things such as tables, chairs, water, etc is made up of matter. The building blocks
of matter are atoms. Therefore, living or non-living thing is composed of atoms.

Materials are composed of very small particles called Atoms. Atom is a Greek word which means
“indivisible.” Atoms are the building blocks of all matter. The Greeks believed that matter can be
broken down into very small invisible particles called atoms. Greek philosophers such as Democritus
and John Dalton put forward the concept of the atom.

Democritus explained the nature of matter. He also proposed that all substances are made up of matter.
In 1808, John Dalton proposed the atomic theory and explained the law of chemical combination.
Many scientists such as J.J Thomson, Gold stein, Rutherford, Bohr among others developed and
proposed several concepts on “atom” by the end of 18th and the early 20th centuries.

All matter have electrical properties. Though an atom is the smallest unit of matter but it retains the
chemical and electrical properties of an element. For example, a silver spoon is made up of silver
atoms with few other constituents. A silver atom obtains its properties from tiny subatomic particles
that it is composed of.
An atom is made up of a positively charged particle at its center called the nucleus. A negatively
charged particle called electrons revolves round the nucleus in different orbits (path) as seen in
figure 1.

Figure 1: An Atom
The Nucleus which is the central part of an atom is made up of positively charged protons and
neutrons with equal mass as the proton but no charges. Hence, the nucleus is positively charged.

4
The weight of an atom known as the atomic weight is the sum of protons and neutrons. The weight
of the electrons is negligible compared to protons and neutrons.

Atomic weight = no. of protons + no. of neutrons

The negatively charged particle the electron is the outer part of the atom. The electrons revolve
round the orbit or centre of the nucleus. The electrons and protons in an atom have equal number
and opposite charges. The number of electrons or protons in an atom is called atomic number. The
attraction between the protons and electrons holds the structure of an atom together.

Atomic number = no. of protons or electrons in an atom

Examples:
1. Lithium atom has 3 electrons revolving around a nucleus of 3 protons and 4 neutrons.
Solution:
Atomic weight of Lithium = number of protons + neutrons
= 3+4
= 7
Atomic number of Lithium = number of protons or electrons
= 3
2. Copper atom has an Atomic weight of 64 and Atomic number of 29. Find the number of
its proton, neutrons and electrons.
Solution:
Atomic number = number of protons or electrons = 29
No. of proton or electron = 29
Atomic weight = number of protons + neutrons
Number of neutrons = 64 – 29 = 35
Assignment 1.1
Find the atomic weight and atomic number showing clearly the protons, neutrons and electrons
of the following elements;
a. Sulphur c. Neon
b. Silicon d Germanium
c. Magnesium

5
1.2 THE STRUCTURE AND COMPOSITION OF AN ATOM
Structure of an atom can be basically divided into two parts:
 an atomic nucleus
 extra nucleus part
The tiny atomic nucleus is the centre of an atom constituting positively charged particles “protons”
and uncharged particles “neutrons.” On the other hand, the extra nucleus part is a much larger region
which is composed of a cloud of negatively charged electron. Electrons revolve around the orbit or
centre of the nucleus. The attraction between the protons and electrons holds the structure of an atom
together.

The atomic nucleus in the structure of the atom is composed of a fixed number of protons and the
proton attracts the same number of electrons thereby making an atom electrically neutral. Ions are
formed by addition or removal of electrons from an atom. The atomic structure of any element can
be built given the atomic weight and number of the element. Elements have different weight and
number hence have different arrangement of particles in their atoms.

1.2.1 Electrons

A British Physicist named J.J Thompson in the year 1897 proposed that an atom constitutes of at least
one negatively charged particle. He named it as “corpuscles” which was later called “electron.” An
electron is represented “e” and it is known to contribute to the negative charge of an atom. The
absolute charge of an electron is the negative charge of 1.6×10-19 coulombs. The relative mass of an
electron is 1/1836, thus the mass of an electron is very small and is considered as 0. The number of
electrons indicates the type of reactions will happen in an atom.

1.2.2 Protons

Rutherford discovered Proton when he conducted the famous gold foil experiment. In 1886 Goldstein
discovered the presence of positively charged rays while conducting an experiment in the discharged
tube by using perforated cathode. The rays were named as anode rays or canal rays. A series of
experiments led to the discovery of protons. Protons are known as the particles that contribute to the
positive charge of the atom. Proton is represented by “p”. The absolute charge of a proton is the
positive charge of 1.6×10-19 coulomb. The mass of a proton is 1.6×10 -24 g and is considered 1 that is
mass of a hydrogen atom. The number of protons indicates what element an atom.

1.2.3 Neutron
6
James Chadwick discovered neutron in 1932. He used scattered particle to calculate the mass of the
neutral particle. The subatomic particle “neutron” is present in an atom’s nucleus. Neutron considered
a neutral particle is represented by “n”. The mass of a neutron is measured to be 1.6 x 10 -24 g. Gram
is however not an appropriate unit for the calculation of such tiny subatomic particles. Therefore they
are alternatively calculated in Dalton or amu (atomic mass unit). Neutron and a proton have a mass
that is nearly 1 amu.

1.2.4 Distribution of Electrons in Different Orbits


Electrons revolve around the nucleus while protons and neutrons are inside the nucleus. This sub-
section describes how electrons revolve and how they are arranged around the nucleus.

Neils Bohr gave the planetary model of an atom. He was the first person to suggest the periodicity in
the properties of the elements. “Bohr atomic model” forms the basis of the electronic structure of an
atom. He described the arrangement of electrons (electronic configuration) in different orbits/shells.

He proposed that electrons are distributed in circular electronic shells (orbits). These electrons revolve
in the orbits around the nucleus from a fixed distance. In this topic, we will learn more about the
electronic configuration of different elements.
i. Bohr Bury Schemes
The distribution of electrons in an atom is called Electronic Configuration. The arrangement of
electrons in each orbit known as “Bohr Bury Schemes” is given by:

Electron in each orbit = 2n2 (1)

Where; n represents the number of the orbit.

This formula helps in the determination of the maximum number of electrons present in an orbit.

Electrons are negatively charged subatomic particles arranged like a cloud of negative charges outside
the nucleus of an atom. The arrangement depends upon of their potential energies in different orbits.
The different energy levels are known as 1, 2, 3, 4….. and the corresponding shells are known as K,
L, M, N and so on as shown in figure 1.2.3

7
Figure 1.2.3: Bohr Bury Scheme

 1st energy level - K shell/orbit


 2nd energy level - L shell/orbit
 3rd energy level - M shell/orbit and so on.

For instance, in figure 1.2.4;

First orbit K 2 × 12 = 2 electrons


Second orbit L 2 × 22 = 8 electrons
Third orbit M 2 × 32 = 18 electrons
ii. Arrangement of Electrons in Orbits

The shells begin from the centre and gradually move outwards. So K shell will always have minimum
energy. Similarly, L shell is a little away from nucleus so it will have higher energy than K shell. The
outermost shell will have maximum energy. Now it is important to understand the distribution and
arrangement of electrons in the atoms of any elements in the different energy levels.

An atom of any element is most stable when it has minimum energy. An atom will first fill the lowest
energy level so as to attain the state of minimum energy. Gradually, the electrons will fill the higher
energy levels. Therefore, electrons will first fill K shell, then L shell, M shell, N shell, and so on.

Figure 1.2.4: Energy Level

iii. Electronic Configuration of Elements

8
According to the postulate of Neils Bohr, “electrons revolve around the centre of an atom (nucleus)
in a predictable pathway named orbits”. The representation of the orbits is done by letters and numbers
such as K, L, M, N, O…. and 1, 2, 3, 4…. respectively. The arrangement and distribution of electrons
in different orbits was given by Bohr and Bury.
The arrangement of electrons in different shells and sub-shells is known as the electronic
configuration of a particular element. The electronic configuration diagram represents an element in
its ground state or stable state. There are a set of rules to remember while distribution of electrons in
different orbits.
 Rule 1: The maximum number of electrons present in a particular shell is calculated by the
formula 2n2, where “n” represents the shell number. For instance, K shell is the first shell and
it can hold up to 2(1)2 = 2 electrons. Similarly, L shell is the second shell and it can hold up to
2(2)2 = 8 electrons. This formula helps to calculate the maximum number of electrons that an
orbit can accommodate.

 Rule 2: The maximum capacity to hold electrons in the outermost shell is 8.

 Rule 3: The electrons will fill the inner shells before the outer shells. First electrons will fill
the K-shell and then L shell and so on. Thus, electronic configuration of elements follows an
ascending order.

Examples of the Electronic Configuration


1. Helium
The atomic number of the element = 2. The total number of electrons present in Helium = 2. The
maximum number of electrons in K shell (1st orbit) = 2. Therefore, shells needed = 1.

Figure 1.2.5: Electronic Configuration of Helium

2. Lithium
9
The atomic number of the element = 3. Lithium has 3 electrons. We can apply rule number 3 to fill
the electrons in different in different orbits. The maximum number of electrons accommodated in K
shell (1st orbit) will be 2. The second orbit will accommodate rest of the electrons. Electronic
configuration of Lithium= 2, 1. Therefore, the total number of shells required = 2.

Figure 1.2.6: Electronic Configuration of Lithium

iv. Uses of the Electronic Configuration


 Electronic Configuration helps to understand the structure of periodic table with respect to each
element.
 It also helps in understanding and explanation of the chemical bonds between the atoms.
 It explains the different properties and peculiar properties of certain elements. For example,
electronic configuration explains the reason for the unique properties of lasers and
semiconductors.
v. Importance of the Electronic Configuration
The electronic configuration is very important and basic part of understanding chemistry. It forms the
basis of the periodic table. Additionally, the stability of any orbital will depend upon the electronic
configuration of that element. It also helps us to understand the arrangement of elements in different
periods and groups.
Assignment 1.2
1. Draw the atomic structure of the following elements with their given electrons in bracket;
a. Sulphur
b. Oxygen
c. Chorine
d. Argon
e. Magnesium
2. Determine the electronic distribution of the first 18 elements.

10
Electronic Distribution of First 18 Elements
Electronic Configuration

S/N Shell Valency


Atomic No. of
Element Symbol
Number Electrons K L M N

1
Hydrogen H 1 1 1 1
2
Helium He 2 2 2 0
3
Lithium Li 3 3 2 1 1
4
Beryllium Be 4 4 2 2 2
5
Boron B 5 5 2 3 3
6
Carbon C 6 6 2 4 4
7
Nitrogen N 7 7 2 5 3
8
Oxygen O 8 8 2 6 2
9
Fluorine F 9 9 2 7 1
10
Neon Ne 10 10 2 8 0
11
Sodium Na 11 11 2 8 1 1
12
Magnesium Mg 12 12 2 8 2 2
13
Aluminum Al 13 13 2 8 3 3
14
Silicon Si 14 14 2 8 4 4
15
Phosphorus P 15 15 2 8 5 3
16
Sulphur S 16 16 2 8 6 2
17
Chlorine Cl 17 17 2 8 7 1
18
Argon Ar 18 18 2 8 8 0

11
iv. Pauli Exclusion Principle: Pauli in 1925 enacted a law which explains the periodic table of
chemical elements. The law states that no two electrons in an electronic system can have the same
set of four quantum numbers, n, l, m1, and m2. This statement is known as the Pauli Exclusion
Principle. Electrons in an atom having the same value of n are said to belong to the same electron
shell. A shell is divided into subshells corresponding to different values of l and identified as s, p,
d, f, which corresponds to l = 0, 1, 2, 3,…respectively.

Table 1.2: Electrons shells and subshells


Shell. . . . . . . K L M N
n.......... 1 2 3 4
l........... 0 0 1 0 1 2 0 1 2 3
Subshell . . . . s s p s p d s p d f
mi . . . . . . . . . 0 0 0, ±1 0 0, ±1 0, ±1, ±2 0 0, ±1 0, ±1, ±2 0, . . . , ±3

𝑁𝑢𝑚𝑏𝑒𝑟 2 2 6 2 6 10 2 6 10 14
𝑜𝑓 }
𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑛𝑠 2 8 18 32

In the exclusion principle as represented in Table 1.2 , there are two states for n = 1 which
1
corresponds to l = 0, mi = 0, and ms = ±2 and are called the 1s states. There are two states
1
corresponding to n = 2, l = 0, m1 = 0, and ms = ±2 called the 2s subshell. Also, six energy levels
1
corresponding to n = 2, l = 1, m1 = -1, 0, or +1, and ms = ±2 which are called the 2p subshells.

Hence, the total number of electrons in the L shell is 2 + 6 = 8, d subshell contains a maximum of
10 electrons, f subshell contains a maximum of 14 electrons.
Using superscript to represent the number of electrons in a subshell, in sodium Na, with an atomic
number Z =12, the electronic configuration is

1s22s22p63s1

This leaves sodium with a single electron in its outer shell, hence is said to be monovalent. Alkali
metals (Li, Na, K, Rb, and Cs) Group 1 of the periodic table are monovalent, hence possess the
same chemical properties. Inert gases He, Ne, A, Kr, and Xe have completely filled shells or
subshells hence, they form very stable configurations.

12
Class Work
Give the electronic configuration of group 4 elements Carbon (6), silicon (14), germanium (32),
tin (50)

1.2.5 The Electron


It has been clearly stated that electrons are negatively charged particle with negligible mass.
Electronics deals with these very tiny particles which has important properties such as:
(i) Charge on an electron, e = 1.602 × 10 −19 coulomb
(ii) Mass of an electron, m = 9.0 × 10 −31 kg
(iii) Radius of an electron, r = 1.9 × 10 −15 metre
The ratio e/m of an electron is 1.77 × 1011 coulombs/kg which means the mass of an electron is
very small as compared to its charge. This property makes an electron very mobile, hence greatly
influenced by electric or magnetic fields.
1.2.6 Energy of an Electron
Electrons due to their small mass can easily move from one level to another. However, the inner
electrons are very strongly bound to the nucleus of the atom, and cannot be easily dislodged; hence
they have the lowest energy. The energy of an electron increases as its distance from the nucleus
increases. Electron in the second orbit has more energy than the electron in the first orbit; electron
in the third orbit has higher energy than electrons in the second orbit.
This electron has the
highest energy
6
3

This
2 electron
has the
1 lowest
energy

Energy levels increase as


the distance from the
nucleus increases

Figure 1.2.6 Energy Level

13
A moving electron has kinetic energy due to its motion as well as potential energy due to the charge
on the nucleus. The total energy of the electron is the sum of these two energies. Electrons in the
last orbit possess very high energy as compared to the electrons in the inner orbits. These last orbit
electrons play an important role in determining the physical, chemical and electrical properties of
a material.

1.2.7 Valence Electrons


The maximum number of electrons that can be in the outer shell of an atom is 8. Valence Electrons
refer to the electrons in the outermost orbit of an atom. The valence electrons determine the
physical, chemical and electrical properties of a material. Valence electrons also determine the
electrical conductivity of materials. These electrons determine whether or not the material is
chemically active; metal or non-metal or, a gas or solid.
Materials are generally classified into conductors, insulators and semi-conductors. Most metals
and semiconductors are crystalline in structure. A crystal is made of a space array of atoms or
molecules called ions which are built up by regular repetition in three dimensions of some
fundamental structural unit. When atoms form crystals, the energy levels of the inner-shell
electrons are not affected significantly by the presence of the neighboring atoms. The levels of the
outer-shell electrons are changed since these electrons are shared by more than one atom in the
crystal.
When the number of valence electrons of an atom is less than 4 the material is usually a metal and
a conductor. Examples are sodium, magnesium and aluminium which have 1, 2 and 3 valence
electrons respectively as shown in figure 1.2.7.

+ + +
11 12 13

a. Sodium b) Magnesium c. Aluminium


Figure 1.2.7

14
When the number of valence electrons of an atom is more than 4, the material is usually a non-
metal and an insulator. Examples are nitrogen, sulphur and neon which have 5, 6 and 8 valence
electrons respectively. When the number of valence electrons of an atom is 4, the material has both
metal and non-metal properties and is usually a semiconductor. Examples are carbon, silicon and
germanium.

1.2.8 Electron Mobility


Electrons mobility refers to how fast an electron can move through a conductor. Electron mobility
is the ability of an electron to move through a metal or semiconductor, in the presence of applied
electric field. Electrons can be dislodged from most atoms, thereby making the atom have a net
positive charge called a positive ion. If a stray electron combines with a normal atom, the atom has
a net negative charge called a negative ion.

The valence electrons of different materials possess different energies. The greater the energy of a
valence electron, the lesser it is bound to the nucleus. In certain substances, particularly metals,
the valence electrons possess so much energy that they are very loosely attached to the nucleus.
These loosely attached valence electrons move at random within the material and are called free
electrons. Free electrons can move through metals, gases and vacuums at high speed. They can
also rest on a surface. Trillions of electrons rest on surfaces or travel through space or matter at
the speed of light.

The free electrons can be easily removed or detached by applying a small amount of external
energy. The motion of a particle of charge q (coulombs), mass m (kilograms), moving with a
velocity v (meters per second) in a field Ɛ (volts per meter) is determined by Newton’s Second
Law. The force f (newtons) on a unit positive charge in an electric field is the electric field intensity
Ɛ at that point.
𝒅𝒗
f = qƐ = 𝒎 𝒅𝒕 (2)

The potential V (volts) of point B with respect to point A is the work done against the field in taking
a unit positive charge from point A to B.
𝑋
V = -∫𝑋 Ɛ𝒅𝒙
0

𝑑𝑉
Where Ɛ now represents the X component of the field; − (3)
𝑑𝑥

15
The negative sign indicated that the electric field is directed from the region of higher potential
to the region of lower potential.
The potential energy U (joules) equals the potential multiplied by the charge q;

U = qV (4)

Considering an electron, q which is the magnitude of the electronic charge is given by – q. The
1
total energy W equals the sum of the potential energy U and the kinetic energy 2 𝑚𝑣 2 remains
1
constant. W = U + 2 𝑚𝑣 2 = constant (5)

The force of attraction between the nucleus and the electron of an atom, using hydrogen as an
example is 𝑞 2 ⁄4𝜋𝜖0 𝑟 2, the electronic charge q is in coulombs, the distance r between the two
particles is in meters, the force is in newtons, and ∈0 is the permittivity of free space.
Applying Newton’s Second Law of motion, the force must be equal to the product of electronic
mass m in kilograms and the acceleration 𝑣 2 ⁄𝑟 toward the nucleus.
𝑞2 𝑚𝑣 2
= (6)
4𝜋𝜖0 𝑟 2 𝑟

Where v is the speed of the electrons in its circular path, in meters per second
The potential energy of the electron at a distance r from the nucleus is −𝑞 2 ⁄4𝜋𝜖0 𝑟 2, and its kinetic
1
energy is 2 𝑚𝑣 2 .

According to the conservation of energy,


1 𝑞2
W = 2 𝑚𝑣 2 − 4𝜋𝜖 𝑟2 (7)
0

Equating (6) and (7)


𝑞2
W=− (8)
8𝜋𝜖0 𝑟

This equation gives the desired relationship between the radius and the energy of the electron,
which shows that the total energy of the electrons is always negative. This shows that the energy
of the electron becomes smaller as it gets closer to the nucleus.

16
1.3 Conductors, Inductors and Semi - Conductors
1.3.1. Conductors
Conductors are materials that allow the flow of electrical current through them. Substances such
as silver, copper and graphite will allow electrical current to pass through them. These substances
have a large number of free electrons. When potential difference is applied across a conductor, the
free electrons move towards the positive terminal of supply, constituting electric current.
In conductors, the valence band is either not fully occupied with electrons, or the filled valence
band overlaps with the empty conduction band. In general, both states occurred at the same time,
the electrons can therefore move inside the partially filled valence band or inside the two
overlapping bands. In conductors there is no band gap between the valence band and conduction
band.

Electron
Energy

Conduction band

Valence band

Conductor
Figure 1.3.: Conduction Band
The atoms that make up the material in conductors especially in metals are loosely bound electrons
which share one or more of the electrons. Their electrons are free to move about, forming an
“electron sea” through which current is conducted. The conductivity of metals is based on the free
electrons (so-called Fermi gas) due to the metal bonding.
The conductivity depends, also on the temperature. If the temperature rises, the metal atoms swing
ever stronger, so that the electrons are constrained in their movements. Consequence, the resistance
increases. The best conductors, gold and silver, are used relatively rare because of the high costs
(gold e.g. for the contacting of the finished chips), aluminum and copper are however alternatives
in the semiconductor technology for the wiring of the individual components of microchips. Salts
can also conduct electricity. Conductivity depends on ions when salt is dissolved in water or
melting to enable the ions to move freely.
1.3.2 Insulators

17
Insulators are substances which have practically no free electrons at ordinary temperature.
Insulators possess no free charge carriers and thus are non-conductive. They poorly or do not allow
the flow of electric currents. Substances like pure water and diamond do not conduct current under
the influence of potential difference.
The atomic bond is based on shared electron pairs of nonmetals. The elements which behave like
non-metals have the desire to catch electrons, thus there are no free electrons which might serve
as charge carriers.
In the solid state, ions are arranged in a grid network. By electrical forces, the particles are held
together. There are no free charge carriers to enable a current flow. Thus substances composed of
ions can be both conductor and insulator. In insulators the valence band is fully occupied with
electrons due to the covalent bonds. The electrons cannot move because they are “locked up”
between the atoms.

Conduction band
Energy

Band gap

Valence band

Insulator
Figure 1.3.2: Insulators

1.3.3 Semiconductors
Semiconductors are solids whose conductivity lies between the conductivity of conductors and
insulators. Due to exchange of electrons - to achieve the noble gas configuration - semiconductors
arrange as lattice structure. Unlike metals, the conductivity increases with increasing temperature.
There are smaller band gaps in semiconductors compared with insulators. At room temperature
electrons from the valence band can be lifted into the conduction band, enabling the electrons to
move freely and act as charge carriers, leaving a hole in the valence band which can be filled by
other electrons in the valence band. Thus one gets wandering holes in the valence band, which can
be viewed as positive charge carriers.
1.4 THE CONCEPTS OF CURRENT AND ELECTRON FLOW

18
When a piece of metal wire like copper, aluminum, silver or iron is in equilibrium, there is no
electric current flowing through it. These materials are conductors—they have free electrons—but
all the free electrons in the wire are distributed uniformly through the piece. Even if they might be
randomly moving around, there is no definite direction that a majority of them is going towards—
no current is flowing. But if we connect two ends of a battery to the two ends of the metal wire, an
electric potential difference is generated by the battery, which disturbs the equilibrium. Electrons,
being negative charges, are pulled by the higher potential point and pushed by the lower potential
point. Their random movement now has a direction—current is flowing through the metal wire.
Figure 1 has a visualization of the electron flow through the wire.
Electrostatic potential (“voltage”) is measured in units of volts (V). Electric current is measured in
units of amperes (A, sometimes “amps” for short). One ampere is defined as approximately 6×1018
electrons going through a conductor in one second.
1.4.1 Voltage Source
Any device that produces voltage output continuously is known as a voltage source. There are two
types of voltage sources, namely;
i. direct voltage source and
ii. alternating voltage source.
i. Direct voltage source. A device which produces direct voltage output continuously is
called a direct voltage source. Common examples are cells and d.c. generators. An
important characteristic of a direct voltage source is that it maintains the same polarity of
the output voltage i.e. positive and negative terminals remain the same. When load
resistance RL is connected across such a source,*current flows from positive terminal to
negative terminal via the load [in figure 1.4].
I

ES RI V RL

Figure 1.4 Direct Voltage Source


𝐸𝑆
Load current, I =
𝑅𝐿 +𝑅𝑖

19
Terminal voltage, V = (𝐸𝑆 − 𝐼𝑅𝑖 ) or IRL

This is called direct current because it has just one direction. The current has one direction
as the source maintains the same polarity of output voltage. The opposition to load current
inside the d.c. source is known as internal resistance Ri. The equivalent circuit of a d.c.
source is the generated e.m.f. ES in series with internal resistance Ri of the source as shown
ii. Alternating voltage source. A device which produces alternating voltage output
continuously is known as alternating voltage source e.g. a.c. generator. An important
characteristic of alternating voltage source is that it periodically reverses the polarity of the
output voltage. When load impedance ZL is connected across such a source, current flows
through the circuit that periodically reverses in direction. This is called alternating current.
Fig. 1.4b
I

ES ZI V ZL

Figure 1.4b Alternating Voltage Source

𝐸𝑆
Load current, I =
𝑍𝐿 +𝑍𝑖

Terminal voltage, V = (𝐸𝑆 − 𝐼𝑍𝑖 ) or IZL

The opposition to load current inside the a.c. source is called its internal impedance Zi. The
equivalent circuit of an a.c. source is the generated e.m.f. Eg (r.m.s.) in series with internal
impedance Zi of the source.
iii. Constant Voltage Source
A voltage source which has very low internal *impedance as compared with external load
impedance is known as a constant voltage source. The output voltage nearly remains the same
when load current changes. It is a power source which provides a constant voltage to a load, despite
changes and variance in load resistance. It provides a steady voltage even if the resistance of the
load varies. This valuable component is required when a circuits needs a steady voltage supply

20
without fluctuations. The graph below shows a voltage constant despite variations in current or
resistance.

Fig 1.4.1: Constant Voltage Source


iv. Fig. 1.4.1 illustrates a constant voltage source. It is a d.c. source of 6 V with internal
resistance R = 0.005 Ω. If the load current varies over a wide range of 1 to 10 A, for
any of these values, the internal drop across Ri (= 0.005 Ω) is less than 0.05 volt.
Therefore, the voltage output of the source is between 5.995 to 5.95 volts. This can be
considered constant voltage compared with the wide variations in load current. i
v. Fig. 1.10 (ii) shows the graph for a constant voltage source. It may be seen that the
output voltage remains constant inspite of the changes in load current. Thus as the load
current changes from 0 to 10 A, the output voltage essentially remains the same (= V).
A constant voltage source is represented as shown in Fig. 1.11

1.4.2 Current
Current is a measure of how many electrons are moving in a conductor. Electrostatic potential is
the measure of the force that got the electrons moving in the first place. The higher the potential
difference between two points, the faster the electrons would move between them, like marbles
rolling down a hill. The relationship between electrostatic potential and current can be further
explained.
How much current will flow when we apply a certain potential difference to the ends of a wire
depends on the type of the wire, or, more properly, on a property of the conductor used. George
Simon Ohm (1789–1854), German scientist, was the first one to formulate the relationship between
voltage and current. In 1826 he put down the law that is known by his name.
Ohm’s Law states that the current through a conductor between two points is directly proportional
to the voltage across the two points.

21
Ohm’s Law relates the voltage drop across a piece of conductor to the amount of current flowing
through it:
V = IR

Where, V is the voltage and I is the current. R is called the resistance. It is the proportionality
constant between the voltage applied to the ends of a conductor and the current flowing through
it. By convention, current is defined as flowing from the positive potential to the negative potential,
or from a high potential point to the low potential point. Note that this is actually the opposite of
the direction that the electrons would be moving in the conductor; electrons, having negative
charge, are attracted to the positive potential.

1.4.3 Resistance
A resistor is represented in electric circuit diagrams. It also displays our convention for the
direction of current in an electrical circuit–from the point of high potential to the point of lower
potential. Resistance is measured in units of ohms (Ω). The resistance of a piece of conductor
depends on the shape, size, and material of the conductor. Basically, the longer the conductor is,
the higher its resistance; and the wider a conductor is, the lower its resistance.

Fig 1.4.3: A Resistor


Note that resistors, as electrical components, do not have a polarity; that is, whichever orientation
you plug them into a circuit does not matter. In electronic circuitry, resistances of thousands and
millions of ohms are used regularly. In the thousands range, the common way to refer to the resistor
size is to use kilo ohms; donated by k or K, or sometimes just K. Thus, a resistance of 1000 = 1k
= 1K; of 25000 = 25k = 25K; of 700 = 0.7K. In the millions range, we use the prefix mega- to
describe the size, meg or M for short: hence 1, 000, 000 = 1M; 35, 000, 000 = 35M.

1.4.4 Resistors Colour Code


There are many different types of Resistor available which can be used in both electrical and
electronic circuits to control the flow of current or to produce a voltage drop in many different

22
ways. But in order to do this the actual resistor needs to have some form of “resistive” or
“resistance” value. Resistors are available in a range of different resistance values from fractions
of an Ohm ( Ω ) to millions of Ohms.
Obviously, it would be impractical to have available resistors of every possible value for example,
1Ω, 2Ω, 3Ω, 4Ω etc, because literally tens of hundreds of thousands, if not tens of millions of
different resistors would need to exist to cover all the possible values. Instead, resistors are
manufactured in what are called “preferred values” with their resistance value printed onto their
body in coloured ink.

Fig 1.4.4: Coloured Bands


The resistance value, tolerance, and wattage rating are generally printed onto the body of the
resistor as numbers or letters when the resistors body is big enough to read the print, such as large
power resistors. But when the resistor is small such as a 1/4 watt carbon or film type, these
specifications must be shown in some other manner as the print would be too small to read.
So to overcome this, small resistors use coloured painted bands to indicate both their resistive
value and their tolerance with the physical size of the resistor indicating its wattage rating. These
coloured painted bands produce a system of identification generally known as a Resistors Colour
Code.
An international and universally accepted resistor colour code scheme was developed many years
ago as a simple and quick way of identifying a resistors ohmic value no matter what its size or
condition. It consists of a set of individual coloured rings or bands in spectral order representing
each digit of the resistors value.
The resistor colour code markings are always read one band at a time starting from the left to the
right, with the larger width tolerance band oriented to the right side indicating its tolerance. By
matching the colour of the first band with its associated number in the digit column of the colour
chart below the first digit is identified and this represents the first digit of the resistance value.
Again, by matching the colour of the second band with its associated number in the digit column
of the colour chart we get the second digit of the resistance value and so on. Then the resistor
colour code is read from left to right as illustrated below:

23
The Standard Resistor Colour Code Chart

Fig 1.4.5: Resistor’s Colour Code


The “left-hand” or the most significant coloured band is the band which is nearest to a connecting
lead with the colour coded bands being read from left-to-right as follows:
Digit, Digit, Multiplier = Colour, Colour x 10 colour in Ohm’s (Ω)
For example, a resistor has the following coloured markings;
Yellow Violet Red = 4 7 2 = 4 7 x 102 = 4700Ω or 4k7 Ohm.
The fourth and fifth bands are used to determine the percentage tolerance of the resistor. Resistor
tolerance is a measure of the resistors variation from the specified resistive value and is a
consequence of the manufacturing process and is expressed as a percentage of its “nominal” or
preferred value.
Typical resistor tolerances for film resistors range from 1% to 10% while carbon resistors have
tolerances up to 20%. Resistors with tolerances lower than 2% are called precision resistors with
the or lower tolerance resistors being more expensive.

24
Most five band resistors are precision resistors with tolerances of either 1% or 2% while most of
the four band resistors have tolerances of 5%, 10% and 20%. The colour code used to denote the
tolerance rating of a resistor is given as:
Brown = 1%, Red = 2%, Gold = 5%, Silver = 10 %
If resistor has no fourth tolerance band then the default tolerance would be at 20%.

1.4.5 Types of Resistor


Resistors are the most commonly used of all the electronic components, it is however taken for
granted. Resistors play a vital role within a circuit. There are many different types of resistor
available ranging from very small surface mount chip resistors to large wire wound power
resistors, for the electronics constructor to choose from.
The principal job of a resistor within an electrical or electronic circuit is to “resist” (hence the
name Resistor), regulate or to set the flow of electrons (current) through them by using the type
of conductive material from which they are composed.
Resistors are “Passive Devices” because they contain no source of power or amplification but only
attenuate or reduce the voltage or current signal passing through them. This attenuation results in
electrical energy being lost in the form of heat as the resistor resists the flow of electrons through
it.
Then a potential difference is required between the two terminals of a resistor for current to flow.
This potential difference balances out the energy lost. When used in DC circuits the potential
difference, also known as a resistors voltage drop, is measured across the terminals as the circuit
current flows through the resistor.
Most types of resistor are linear devices that produce a voltage drop across themselves when an
electrical current flows through them because they obey Ohm’s Law, and different values of
resistance produces different values of current or voltage. This can be very useful in Electronic
circuits by controlling or reducing either the current flow or voltage produced across them we can
produce a voltage-to-current and current-to-voltage converter.
There are many thousands of different types of resistor and are produced in a variety of forms
because their particular characteristics and accuracy suit certain areas of application, such as High
Stability, High Voltage, High Current etc, or are used as general purpose resistors where their
characteristics are less of a problem.

25
Some of the common characteristics associated with the humble resistor are; Temperature
Coefficient, Voltage Coefficient, Noise, Frequency Response, Power as well as a resistors
Temperature Rating, Physical Size and Reliability.
In all Electrical and Electronic circuit diagrams and schematics, the most commonly used symbol
for a fixed value resistor is that of a “zig-zag” type line with the value of its resistance given in
Ohms, Ω. Resistors have fixed resistance values from less than one ohm, ( <1Ω ) to well over tens
of millions of ohms, ( >10MΩ ) in value.
Fixed resistors have only one single value of resistance, for example 100Ω, but variable resistors
(potentiometers) can provide an infinite number of resistance values between zero and their
maximum value.

Fig 1.4.6: Standard Resistor Symbols

The symbol commonly used in schematic and electrical drawings for a Resistor can either be a
“zig-zag” type line or a rectangular box.
All modern fixed value resistors can be classified into four broad groups:
 Carbon Composition Resistor – Made of carbon dust or graphite paste, low wattage values
 Film or Cermet Resistor – Made from conductive metal oxide paste, very low wattage values
 Wire-wound Resistor – Metallic bodies for heat sink mounting, very high wattage ratings
 Semiconductor Resistor – High frequency/precision surface mount thin film technology
There are a large variety of fixed and variable resistor types with different construction styles
available for each group, with each one having its own particular characteristics, advantages and
disadvantages compared to the others.

1.4.6 Composition Types of Resistor


i. Carbon Resistors are the most common type of Composition Resistors. Carbon resistors are
a cheap general purpose resistor used in electrical and electronic circuits. Their resistive element
is manufactured from a mixture of finely ground carbon dust or graphite (similar to pencil lead)
and a non-conducting ceramic (clay) powder to bind it all together.

26
Fig 1.4.7: Carbon Resistor
The ratio of carbon dust to ceramic (conductor to insulator) determines the overall resistive value
of the mixture and the higher the ratio of carbon, the lower the overall resistance. The mixture is
moulded into a cylindrical shape with metal wires or leads are attached to each end to provide the
electrical connection as shown, before being coated with an outer insulating material and colour
coded markings to denote its resistive value.
ii. The Carbon Composite Resistor is a low to medium type power resistor which has a low
inductance making them ideal for high frequency applications but they can also suffer from
noise and stability when hot. Carbon composite resistors are generally prefixed with a “CR”
notation (eg, CR10kΩ ) and are available in E6 ( ± 20% tolerance (accuracy) ), E12 ( ± 10%
tolerance) and E24 ( ± 5% tolerance) packages with power ratings from 0.250 or 1/4 of a Watt
up to 5 Watts.

Fig 1.4.8: The Carbon Composite Resistor


Carbon composite resistor types are very cheap to make and are therefore commonly used in
electrical circuits. However, due to their manufacturing process carbon type resistors have very
large tolerances so for more precision and high value resistances, film type resistors are used
instead.

iii. Film Type Resistors

27
The generic term “Film Resistor” consist of Metal Film, Carbon Film and Metal Oxide Film
resistor types, which are generally made by depositing pure metals, such as nickel, or an oxide
film, such as tin-oxide, onto an insulating ceramic rod or substrate.
The resistive value of the resistor is controlled by increasing the desired thickness of the deposited
film giving them the names of either “thick-film resistors” or “thin-film resistors”.
Once deposited, a laser is used to cut a high precision spiral helix groove type pattern into this
film. The cutting of the film has the effect of increasing the conductive or resistive path, a bit like
taking a long length of straight wire and forming it into a coil.
This method of manufacture allows for much closer tolerance resistors (1% or less) as compared
to the simpler carbon composition types. The tolerance of a resistor is the difference between the
preferred value (i.e, 100 ohms) and its actual manufactured value i.e, 103.6 ohms, and is expressed
as a percentage, for example 5%, 10% etc, and in our example the actual tolerance is 3.6%. Film
type resistors also achieve a much higher maximum ohmic value compared to other types and
values in excess of 10MΩ (10 Million Ohms) are available.

Fig 1.4.9: Film Resistor

iv. Metal Film Resistors


Metal Film Resistors have much better temperature stability than their carbon equivalents, lower
noise and are generally better for high frequency or radio frequency applications. Metal Oxide
Resistors have better high surge current capability with a much higher temperature rating than the
equivalent metal film resistors.
Another type of film resistor commonly known as a Thick Film Resistor is manufactured by
depositing a much thicker conductive paste of CERamic and METal, called Cermet, onto an
alumina ceramic substrate. Cermet resistors have similar properties to metal film resistors and are
generally used for making small surface mount chip type resistors, multi-resistor networks in one
28
package for pcb’s and high frequency resistors. They have good temperature stability, low noise,
and good voltage ratings but low surge current properties.
Metal Film Resistors are prefixed with a “MFR” notation (eg, MFR100kΩ) and a CF for Carbon
Film types. Metal film resistors are available in E24 (±5% & ±2% tolerances), E96 (±1% tolerance)
and E192 (±0.5%, ±0.25% & ±0.1% tolerances) packages with power ratings of 0.05 (1/20th) of a
Watt up to 1/2 Watt. Generally speaking Film resistors and especially metal film resistors are
precision low power components.
v. Wire-wound Types of Resistor
Another type of resistor, called a Wire-wound Resistor, is made by winding a thin metal alloy
wire (Nichrome) or similar wire onto an insulating ceramic former in the form of a spiral helix
similar to the film resistor above.
These types of resistor are generally only available in very low ohmic high precision values (from
0.01Ω to 100kΩ) due to the gauge of the wire and number of turns possible on the former making
them ideal for use in measuring circuits and Wheatstone bridge type applications.
They are also able to handle much higher electrical currents than other resistors of the same ohmic
value with power ratings in excess of 300 Watts. These high power resistors are moulded or
pressed into an aluminium heat sink body with fins attached to increase their overall surface area
to promote heat loss and cooling.
These special types of resistor are called “Chassis Mounted Resistors” because they are designed
to be physically mounted onto heat sinks or metal plates to further dissipate the generated heat.
The mounting of the resistor onto a heat sink increases their current carrying capabilities even
further.
Another type of wire-wound resistor is the Power Wire-wound Resistor. These are high
temperature, high power non-inductive resistor types generally coated with a vitreous or glass
epoxy enamel for use in resistance banks or DC motor/servo control and dynamic braking
applications. They can even be used as low wattage space or cabinet heaters.
The non-inductive resistance wire is wound around a ceramic or porcelain tube covered with mica
to prevent the alloy wires from moving when hot. Wire-wound resistors are available in a variety
of resistance and power ratings with one main use of power wire-wound resistor is in the electrical
heating elements of an electric fire which converts the electrical current flowing through it into
heat with each element dissipating up to 1000 Watts, (1kW) of energy.

29
Because the wire of standard wire wound resistors is wound into a coil inside the resistors body, it
acts like an inductor causing them to have inductance as well as resistance. This affects the way
the resistor behaves in AC circuits by producing a phase shift at high frequencies especially in the
larger size resistors. The length of the actual resistance path in the resistor and the leads contributes
inductance in series with the “apparent” DC resistance resulting in an overall impedance path of Z
Ohms.
Impedance ( Z ) is the combined effect of resistance ( R ) and inductance ( X ), measured in ohms
and for a series AC circuit is given as, Z2 = R2 + X2.
When used in AC circuits this inductance value changes with frequency (inductive reactance,
XL = 2πƒL) and therefore, the overall value of the resistor changes. Inductive reactance increases
with frequency but is zero at DC (zero frequency). Then, wire wound resistors must not be
designed or used in AC or amplifier type circuits where the frequency across the resistor changes.
However, special non-inductive wire wound resistors are also available.

Fig 1.4.10: Wire wound Resistor


Wire wound resistor types are prefixed with a “WH” or “W” notation (eg WH10Ω) and are
available in the WH aluminium clad package (±1%, ±2%, ±5% and ±10% tolerance) or the W
vitreous enamelled package (±1%, ±2% and ±5% tolerance) with power ratings from 1W to 300W
or more.

30
Table 1.4.2: SI Unit for Electrical Quantities

Electrical Parameter Measuring Unit Symbol Description


Voltage Volts V or E Unit of Electrical Potential
V=I*R
Current Ampere I or i Unit of Electrical Current
I=V/R
Resistance Ohms R or Ω Unit of DC Resistance
R=V/I
Conductance Siemens G or ʊ Reciprocal of Resistance
G=1/R
Capacitance Farad C Unit of Capacitance
C=Q*V
Charge Coulomb C Unit of Electrical Charge
Q=C/V
Inductance Henry L or H Unit of Inductance
Vi = Li[di(t) / dt]
Power Watts W Unit of Power
P = V * I or I2*R
Impedance Ohms Z Unit of AC Resistance
Z2 = R2 – X2
Frequency Hertz Hz Unit of Frequency
f=1/T

31
Table 1.4.3: Multiples and Sub – multiples of Electrical Quantities
Prefix Symbol Multiplication Factor
exa E 1018 = 1 000 000 000 000 000 000
peta P 1015 = 1 000 000 000 000 000
tera T 1012 = 1 000 000 000 000
giga G 109 = 1 000 000 000
mega M 106 = 1 000 000
kilo k 103 = 1 000
hector h 102 = 100
deca da 101 = 10
deci d 10-1 = 0.1
centi c 10-2 = 0.01
milli m 10-3 = 0.001
micro µ 10-6 = 0.000 001
nano n 10-9 = 0.000 000 001
pico p 10-12 = 0.000 000 000 001
femto f 10-15 = 0.000 000 000 000 001
atto a 10-18 = 0.000 000 000 000 000 001

32
MODULE TWO

2.1 D.C. CURRENT


Direct current (DC) is the unidirectional flow of electric charge. The electric charge flows in only
one direction, steadily or in pulse. A battery is a good example of a DC power supply.
DC circuit, electrons emerge from the negative, or minus, pole and move towards the positive, or
plus, pole. Nevertheless, physicists define DC as traveling from plus to minus. Virtually all
electronic and computer hardware needs DC to function. Most solid-state equipment requires
between 1.5 and 13.5 volts.

Fig 2.1: Direct Current

2.2 The Analogy between Current Flow and Water Flow


When describing voltage, current, and resistance, a common analogy is a water tank. In this
analogy, charge is represented by the water amount, voltage is represented by the water pressure,
and current is represented by the water flow. So for this analogy, remember:
Water = Charge
Pressure = Voltage
Flow = Current
The pressure at the end of the hose can represent voltage. The water in the tank represents charge.
The more water in the tank, the higher the charge, the more pressure is measured at the end of the
hose. We can think of this tank as a battery, a place where we store a certain amount of energy and
then release it. If we drain our tank a certain amount, the pressure created at the end of the hose
goes down. There is also a decrease in the amount of water that will flow through the hose. Less
pressure means less water is flowing, which brings us to current.

33
Consider a water tank at a certain height above the ground. At the bottom of this tank there is a
hose.

Fig 2.2: Analogy between Current Flow and Water Flow

2.4 Constant Current Source


A voltage source that has very high internal impedance as compared with external load impedance
is considered as a constant current source. The load current nearly remains the same when the
output voltage changes.
A d.c. source of 1000 V with internal resistance R = 900kΩ. Here, load R varies over 3: 1 range
from 50k Ω to 150kΩ. Over this variation of load R L, the circuit current I is essentially constant
at 1.05 to 0.95mA or approximately 1 mA. It may be noted that output voltage V varies
approximately in the same 3: 1 range as RL, although load current essentially remains constant at
1mA.
The example of a constant current source is found in vacuum tube circuits where the tube acts as
a generator having internal resistance as high as 1 MΩ. The following points may be noted
regarding the constant current source:
(i) Due to high internal resistance of the source, the load current remains essentially constant as
the load R is varied.
(ii) The output voltage varies approximately in the same range as RL, although current remains
constant.
(iii) The output voltage V is much less than the generated voltage E because of high IR drop.

34
Example 1
1. A d.c. source generating 500 V has an internal resistance of 1000 Ω. Find the load current if
load resistance is
i. 10 Ω
ii. 50 Ω and
iii. 100 Ω.

2.4 OHM’S LAW


The relationship between Voltage, Current and Resistance in any DC electrical circuit was firstly
discovered by the German physicist Georg Ohm.
Georg Ohm a German physicist found that, at a constant temperature, the electrical current flowing
through a fixed linear resistance is directly proportional to the voltage applied across it, and also
inversely proportional to the resistance. This relationship between the Voltage, Current and
Resistance forms the basis of Ohms Law and is shown below.
Ohm's law states that the current through a conductor between two points is directly proportional
to the voltage across the two points.
𝑉
𝐼=
𝑅
Where I is the current through the conductor in units of amperes, V is the voltage measured across
the conductor in units of volts, and R the constant of proportionality is resistance of the conductor
in units of ohms. More specifically, Ohm's law states that the R in this relation is constant,
independent of the current. Ohm’s Law Triangle is given as;

Fig 2.3: Ohm’s Law Triangle


Then by using Ohms Law we can see that a voltage of 1V applied to a resistor of 1Ω will cause a
current of 1A to flow and the greater the resistance value, the less current that will flow for a given

35
applied voltage. Any Electrical device or component that obeys “Ohms Law” that is, the current
flowing through it is proportional to the voltage across it ( I α V ), such as resistors or cables, are
said to be “Ohmic” or "linear" in nature, and devices that do not, such as transistors or diodes,
are said to be “Non-ohmic” devices.
Ohm's Law can be used to solve simple circuits. A complete circuit is one which is a closed loop.
It contains at least one source of voltage (thus providing an increase of potential energy), and at
least one potential drop i.e., a place where potential energy decreases. The sum of the voltages
around a complete circuit is zero.
An increase of potential energy in a circuit causes a charge to move from a lower to a higher
potential (ie. voltage). Note the difference between potential energy and potential. Because of the
electrostatic force, which tries to move a positive charge from a higher to a lower potential, there
must be another 'force' to move charge from a lower potential to a higher inside the battery. This
so-called force is called the electromotive force, or emf. The SI unit for the emf, E is a volt. The
symbol E, represents the emf.
A decrease of potential energy can occur by various means. For example, heat lost in a circuit due
to some electrical resistance could be one source of energy drop. Because energy is conserved, the
potential difference across an emf must be equal to the potential difference across the rest of the
circuit. That is, Ohm's Law will be satisfied:

=IR

2.5 Practice Problems


1. A circuit with a source voltage of 12V DC and a circuit resistance of 150Ω. Solve for the
unknown using Ohm’s Law.
2.

36
3.

4. I = 2A, R = 55ohms. What is the value of E?


5. A Resistance of 15Ω carries a current of 5Amps. Calculate the voltage developed across the
resistor.
6. Calculate the resistance of an iron filament if it operates on 230V supply and draws a current
equal to 2Amp.
7. 1. A toaster oven is plugged into an outlet that provides a voltage difference of 120 V. What
power does the oven use if the current is 10A?
8. 2. A VCR that is not playing still uses 10.0 W of power. What is the current if the VCR is
plugged into a 120 V electric outlet?
9. 3. A flashlight bulb uses 2.4 W of power when the current in the bulb is 0.8 A. What is the
voltage difference?
10. 4. A refrigerator operates on average for 10.0 h a day. If the power rating of the refrigerator is
700 W, how much electrical energy does the refrigerator use in 1 day? (make sure to convert
to kW)

2.6 Electrical Power in Circuits


Electrical Power, (P) in a circuit is the rate at which energy is absorbed or produced within a circuit.
A source of energy such as a voltage will produce or deliver power while the connected load
absorbs it. Light bulbs and heaters for example, absorb electrical power and convert it into either
heat, or light, or both. The higher their value or rating in watts the more electrical power they are
likely to consume.
The quantity symbol for power is P and is the product of voltage multiplied by the current with the
unit of measurement being the Watt (W). Prefixes are used to denote the various multiples or sub-
multiples of a watt, such as: milliwatts (mW = 10-3W) or kilowatts (kW = 103W).
Then by using Ohm’s law and substituting for the values of V, I and R the formula for electrical
power can be found as:

37
Power (P) P (watts) = V (volts) x I (amps)
P=VxI

Also: P (watts) = V2 (volts) ÷ R (Ω)


P = V2 ÷ R

Also: P (watts) = I2 (amps) x R (Ω)


P = I2 x R

Fig 2.4: Power Triangle

2.6.1 Electrical Power Rating


Electrical components are given a “power rating” in watts that indicates the maximum rate at which
the component converts the electrical power into other forms of energy such as heat, light or
motion. For example, a 1/4W resistor, a 100W light bulb etc.
Electrical devices convert one form of power into another. So for example, an electrical motor will
covert electrical energy into a mechanical force, while an electrical generator converts mechanical
force into electrical energy. A light bulb converts electrical energy into both light and heat.
Also, we now know that the unit of power is the WATT, but some electrical devices such as electric
motors have a power rating in the old measurement of “Horsepower” or hp. The relationship
between horsepower and watts is given as: 1hp = 746W. So for example, a two-horsepower motor
has a rating of 1492W, (2 x 746) or 1.5kW.

2.7 Resistivity
Electrical resistivity (also known as specific electrical resistance or volume resistivity) is a
fundamental property of a material that quantifies how strongly that material opposes the flow of

38
electric current. A low resistivity indicates a material that readily allows the flow of electric
current. Resistivity is commonly represented by the Greek letter ρ (rho). The SI unit of electrical
resistivity is the ohm-metre (Ω⋅m). Resistivity of materials is the resistance to the flow of an
electric current with some materials resisting the current flow more than others
Ohms Law states that when a voltage (V) source is applied between two points in a circuit, an
electrical current (I) will flow between them encouraged by the presence of the potential difference
between these two points. The amount of electrical current which flows is restricted by the amount
of resistance (R) present. In other words, the voltage encourages the current to flow (the movement
of charge), but it is resistance that discourages it.
We always measure electrical resistance in Ohms, where Ohms is denoted by the Greek letter
Omega, Ω. So for example: 50Ω, 10kΩ or 4.7MΩ, etc. Conductors (e.g. wires and cables) generally
have very low values of resistance (less than 0.1Ω) and thus we can neglect them as we assume in
circuit analysis calculations that wires have zero resistance. Insulators (e.g. plastic or air) on the
other hand generally have very high values of resistance (greater than 50MΩ), therefore we can
ignore them also for circuit analysis as their value is too high.
The electrical resistance between two points can depend on many factors such;
i. as the conductors length,
ii. its cross-sectional area,
iii. the temperature,
iv. the actual material from which it is made.
For example, let’s assume we have a piece of wire (a conductor) that has a length L, a cross-
sectional area A and a resistance R as shown.

Fig 2.5: Resistivity of L Length Conductor

The electrical resistance, R of this simple conductor is a function of its length, L and the conductors
area, A. Ohms law tells us that for a given resistance R, the current flowing through the conductor

39
is proportional to the applied voltage as I = V/R. Now suppose we connect two identical conductors
together in a series combination as shown.

Fig 2.6: Resistivity of 2L Length Conductor

By connecting the two conductors together in a series combination, that is end to end, we have
effectively doubled the total length of the conductor (2L), while the cross-sectional area A, remains
exactly the same as before. But as well as doubling the length, we have also doubled the total
resistance of the conductor, giving 2R as: 1R + 1R = 2R.
Therefore we can see that the resistance of the conductor is proportional to its length, that is:
R ∝ L. In other words, we would expect the electrical resistance of a conductor (or wire) to be
proportionally greater the longer it is.
Note also that by doubling the length and therefore the resistance of the conductor (2R), to force
the same current, i to flow through the conductor as before, we need to double (increase) the
applied voltage as now I = (2V)/(2R). Suppose we connect the two identical conductors together
in parallel combination, we have effectively doubled the total area giving 2A, while the
conductor’s length, L remains the same as the original single conductor. But as well as doubling
the area, by connecting the two conductors together in parallel we have effectively halved the total
resistance of the conductor, giving 1/2R as now each half of the current flows through each
conductor branch.
Thus the resistance of the conductor is inversely proportional to its area, that is: R 1/∝ A, or
R ∝ 1/A. In other words, we would expect the electrical resistance of a conductor (or wire) to be
proportionally less the greater is its cross-sectional area. Also by doubling the area and therefore
halving the total resistance of the conductor branch (1/2R), for the same current, i to flow through
the parallel conductor branch as before we only need half (decrease) the applied voltage as now I
= (1/2V)/(1/2R).
So hopefully we can see that the resistance of a conductor is directly proportional to the length (L)
of the conductor, that is: R ∝ L, and inversely proportional to its area (A), R ∝ 1/A. Thus we can
correctly say that resistance is:
40
𝒍
𝑹= 𝝆
𝑨
Where: R is the resistance in ohms (Ω), L is the length in metres (m), A is the area in square metres
(m2), and where the proportional constant ρ (the Greek letter “rho”) is known as Resistivity.
The electrical resistivity of a particular conductor material is a measure of how strongly the
material opposes the flow of electric current through it. This resistivity factor, sometimes called
its “specific electrical resistance”, enables the resistance of different types of conductors to be
compared to one another at a specified temperature according to their physical properties without
regards to their lengths or cross-sectional areas. Thus the higher the resistivity value of ρ the more
resistance and vice versa.
For example, the resistivity of a good conductor such as copper is on the order of 1.72 x 10-8 ohm
metre (or 17.2 nΩm), whereas the resistivity of a poor conductor (insulator) such as air can be well
over 1.5 x 1014 or 150 trillion Ωm.
Materials such as copper and aluminium are known for their low levels of resistivity thus allowing
electrical current to easily flow through them making these materials ideal for making electrical
wires and cables. Silver and gold have much low resistivity values, but for obvious reasons are
more expensive to turn into electrical wires.
Then the factors which affect the resistance (R) of a conductor in ohms can be listed as:
 The resistivity (ρ) of the material from which the conductor is made.
 The total length (L) of the conductor.
 The cross-sectional area (A) of the conductor.
 The temperature of the conductor.

Practical Problems
1. Calculate the total DC resistance of a 100 metre roll of 2.5mm2 copper wire if the resistivity
of copper at 20oC is 1.72 x 10-8 Ω metre. Data given: resistivity of copper at 20oC is 1.72 x 10-
8
, coil length L = 100m, the cross-sectional area of the conductor is 2.5mm2 giving an area of:
A = 2.5 x 10-6 metres2.
2. A 20 metre length of cable has a cross-sectional area of 1mm2 and a resistance of 5 ohms.
Calculate the conductivity of the cable. Data given: DC resistance, R = 5 ohms, cable length,
L = 20m, and the cross-sectional area of the conductor is 1mm2 giving an area of: A = 1 x 10-
6
metres2.
41
2.8 Electrical Conductivity
Electrical conductivity or specific conductance is the reciprocal of electrical resistivity, and
measures a material's ability to conduct an electric current. It is commonly represented by the
Greek letter σ (sigma), but κ (kappa) (especially in electrical engineering) or γ (gamma) are also
occasionally used. Its SI unit is siemens per metre (S/m).
While both the electrical resistance (R) and resistivity (or specific resistance) ρ, are a function of
the physical nature of the material being used, and of its physical shape and size expressed by its
length (L), and its sectional area (A), Conductivity, or specific conductance relates to the ease at
which electric current can flow through a material.
Conductance (G) is the reciprocal of resistance (1/R) with the unit of conductance being the
siemens (S) and is given the upside down ohms symbol mho, ℧. Thus when a conductor has a
conductance of 1 siemens (1S) it has a resistance is 1 ohm (1Ω). So if its resistance is doubled, the
conductance halves, and vice-versa as: siemens = 1/ohms, or ohms = 1/siemens.
While a conductor’s resistance gives the amount of opposition it offers to the flow of electric
current, the conductance of a conductor indicates the ease by which it allows electric current to
flow. So metals such as copper, aluminium or silver have very large values of conductance
meaning that they are good conductors.
Conductivity, σ (Greek letter sigma), is the reciprocal of the resistivity. That is 1/ρ and is measured
in siemens per metre (S/m). Since electrical conductivity σ = 1/ρ, the previous expression for
electrical resistance, R can be rewritten as:
𝒍
𝑹= 𝝆
𝑨
Electrical Resistance as a Function of Conductivity
𝟏
𝝈=
𝝆
𝒍
𝑹= Ω
𝝈𝑨
2.9 Series and Parallel Circuits
Resistors can be connected together in various series and parallel combinations to form resistor
networks which can act as voltage droppers, voltage dividers or current limiters within a circuit.
Individual resistors can be connected together in either a series connection, a parallel connection
or combinations of both series and parallel, to produce more complex resistor networks whose

42
equivalent resistance is the mathematical combination of the individual resistors connected
together.
A resistor is not only a fundamental electronic component that can be used to convert a voltage to
a current or a current to a voltage, but by correctly adjusting its value a different weighting can be
placed onto the converted current and/or the voltage allowing it to be used in voltage reference
circuits and applications.
Resistors in series or complicated resistor networks can be replaced by one single equivalent
resistor, REQ or impedance, ZEQ and no matter what the combination or complexity of the resistor
network is, all resistors obey the same basic rules as defined by Ohm’s Law and Kirchhoff’s Circuit
Laws.
2.9.1 Resistors in Series
Resistors are said to be connected in “Series”, when they are daisy chained together in a single
line. Since all the current flowing through the first resistor has no other way to go it must also pass
through the second resistor and the third and so on. Then, resistors in series have a Common
Current flowing through them as the current that flows through one resistor must also flow
through the others as it can only take one path.
Then the amount of current that flows through a set of resistors in series will be the same at all
points in a series resistor network. For example:
𝐼𝑅1 = 𝐼𝑅2 = 𝐼𝑅3 = 𝐼𝑅4 = 1𝑚𝐴
In the following example the resistors R1, R2 and R3 are all connected together in series between points
A and B with a common current, I flowing through them.

Fig 2.9.1: Series Resistor Circuit


As the resistors are connected together in series the same current passes through each resistor in
the chain and the total resistance, RT of the circuit must be equal to the sum of all the individual
resistors added together. That is
𝑅𝑇 = 𝑅1 + 𝑅2 + 𝑅3
The individual resistors in series in fig 2.9.1 can be added to give an equivalent resistance, REQ;
REQ = R1 + R2 + R3 = 1kΩ + 2kΩ + 6kΩ = 9kΩ
43
All three individual resistors above can be replaced with just one single “equivalent” resistor which
will have a value of 9kΩ.
Where four, five or even more resistors are all connected together in a series circuit, the total or
equivalent resistance of the circuit, RT would still be the sum of all the individual resistors
connected together and the more resistors added to the series, the greater the equivalent resistance
(no matter what their value).
This total resistance is generally known as the Equivalent Resistance and can be defined as; “a
single value of resistance that can replace any number of resistors in series without altering the
values of the current or the voltage in the circuit“. Then the equation given for calculating total
resistance of the circuit when connecting together resistors in series is given as:
Series Resistor Equation
𝑹𝑻𝒐𝒕 = 𝑹𝟏 + 𝑹𝟐 + 𝑹𝟑 +. . . 𝑹𝒏
It is important to note that the total resistance (RT) of any two or more resistors connected together
in series will always be GREATER than the value of the largest resistor in the chain. In our
example above RT = 9kΩ where as the largest value resistor is only 6kΩ.

2.9.1.1 Series Resistor Voltage


The voltage across each resistor connected in series follows different rules to that of the series
current. We know from the above circuit that the total supply voltage across the resistors is equal
to the sum of the potential differences across R1, R2 and R3,
VAB = VR1 + VR2 + VR3 = 9V.
Using Ohm’s Law, the voltage across the individual resistors can be calculated as:
Voltage across R1 = IR1 = 1mA x 1kΩ = 1V
Voltage across R2 = IR2 = 1mA x 2kΩ = 2V
Voltage across R3 = IR3 = 1mA x 6kΩ = 6V

44
giving a total voltage VAB of ( 1V + 2V + 6V ) = 9V which is equal to the value of the supply
voltage. Then the sum of the potential differences across the resistors is equal to the total potential
difference across the combination and in our example this is 9V.
The equation given for calculating the total voltage in a series circuit which is the sum of all the
individual voltages added together is given as:
𝑉𝑇𝑜𝑡𝑎𝑙 = 𝑉𝑅1 + 𝑉𝑅2 + 𝑉𝑅3 +. . . +𝑉𝑁
Then series resistor networks can also be thought of as “voltage dividers” and a series resistor
circuit having N resistive components will have N-different voltages across it while maintaining a
common current.
By using Ohm’s Law, either the voltage, current or resistance of any series connected circuit can
easily be found and resistor of a series circuit can be interchanged without affecting the total
resistance, current, or power to each resistor.
Example 1:
Using Ohms Law, calculate the equivalent series resistance, the series current, voltage drop and
power for each resistor in the following resistors in series circuit.

All the data can be found by using Ohm’s Law and present in tabular form.

Resistance Current Voltage Power

R1 = 10Ω I1 = 200mA V1 = 2V P1 = 0.4W

R2 = 20Ω I2 = 200mA V2 = 4V P2 = 0.8W

R3 = 30Ω I3 = 200mA V3 = 6V P3 = 1.2W

45
RT = 60Ω IT = 200mA VS = 12V PT = 2.4W

Then for the circuit above, RT = 60Ω, IT = 200mA, VS = 12V and PT = 2.4W

2.9.1.2 The Voltage Divider Circuit


We can see from the above example, that although the supply voltage is given as 12 volts, different
voltages, or voltage drops, appear across each resistor within the series network. Connecting
resistors in series like this across a single DC supply has one major advantage; different voltages
appear across each resistor producing a very handy circuit called a Voltage Divider Network.
This simple circuit splits the supply voltage proportionally across each resistor in the series chain
with the amount of voltage drop being determined by the resistors value and as we now know, the
current through a series resistor circuit is common to all resistors. So a larger resistance will have
a larger voltage drop across it, while a smaller resistance will have a smaller voltage drop across
it.
The series resistive circuit shown above forms a simple voltage divider network was three voltages
2V, 4V and 6V are produced from a single 12V supply. Kirchhoff’s Voltage Law states that “the
supply voltage in a closed circuit is equal to the sum of all the voltage drops (I*R) around the
circuit” and this can be used to good effect.
The Voltage Division Rule, allows us to use the effects of resistance proportionality to calculate
the potential difference across each resistance regardless of the current flowing through the series
circuit. A typical “voltage divider circuit” is shown below.

Fig 2.9.2: Voltage Divider Network

The circuit shown consists of just two resistors, R1 and R2 connected together in series across the
supply voltage Vin. One side of the power supply voltage is connected to resistor, R1, and the
voltage output, Vout is taken from across resistor R2. The value of this output voltage is given by
the corresponding formula.

46
If more resistors are connected in series to the circuit then different voltages will appear across
each resistor in turn with regards to their individual resistance R (Ohms Law I*R) values providing
different but smaller voltage points from one single supply.
So if we had three or more resistances in the series chain, we can still use our now familiar potential
divider formula to find the voltage drop across each one. Consider the circuit below.

The potential divider circuit above shows four resistances connected together in series. The
voltage drop across points A and B can be calculated using the potential divider formula as follows:
𝑅3
𝑉𝐴𝐵 = 𝑉𝑅3 = 𝑉𝑆 ∗
𝑅1 + 𝑅2 + 𝑅3 + 𝑅4
30
𝑉𝐴𝐵 = 𝑉𝑅3 = 10 ∗ = 10 * 0.3 = 3V
10+20+30+40
We can also apply the same idea to a group of resistors in the series chain. For example if we
wanted to find the voltage drop across both R2 and R3 together we would substitute their values
in the top numerator of the formula and in this case the resulting answer would give us 5 volts (2V
+ 3V).
In this very simple example the voltages work out very neatly as the voltage drop across a resistor
is proportional to the total resistance, and as the total resistance, (RT) in this example is equal to
100Ω or 100%, resistor R1 is 10% of RT, so 10% of the source voltage VS will appear across it,
20% of VS across resistor R2, 30% across resistor R3, and 40% of the supply voltage VS across
resistor R4. Application of Kirchhoff’s voltage law (KVL) around the closed loop path confirms
this.
Now let’s suppose that we want to use our two resistor potential divider circuit above to produce
a smaller voltage from a larger supply voltage to power an external electronic circuit. Suppose we
have a 12V DC supply and our circuit which has an impedance of 50Ω requires only a 6V supply,
half the voltage.

47
Connecting two equal value resistors, of say 50Ω each, together as a potential divider network
across the 12V will do this very nicely until we connect the load circuit to the network. This is
because the loading effect of resistor RL connected in parallel across R2 changes the ratio of the
two series resistances altering their voltage drop and this is demonstrated below.
Example 2
Calculate the voltage drops across X and Y
a) Without RL connected
b) With RL connected

As you can see from above, the output voltage Vout without the load resistor connected gives us
the required output voltage of 6V but the same output voltage at Vout when the load is connected
drops to only 4V, (Resistors in Parallel).
Then we can see that a loaded voltage divider network changes its output voltage as a result of this
loading effect, since the output voltage Vout is determined by the ratio of R1 to R2. However, as the
load resistance, RL increases towards infinity (∞) this loading effect reduces and the voltage ratio
of Vout/Vs becomes unaffected by the addition of the load on the output. Then the higher the load
impedance the less is the loading effect on the output.
The effect of reducing a signal or voltage level is known as Attenuation so care must be taken
when using a voltage divider network. This loading effect could be compensated for by using a
48
potentiometer instead of fixed value resistors and adjusted accordingly. This method also
compensates the potential divider for varying tolerances in the resistors construction.
A variable resistor, potentiometer or pot as it is more commonly called, is a good example of a
multi-resistor voltage divider within a single package as it can be thought of as thousands of mini-
resistors in series. Here a fixed voltage is applied across the two outer fixed connections and the
variable output voltage is taken from the wiper terminal. Multi-turn pots allow for a more accurate
output voltage control.
The Voltage Divider Circuit is the simplest way of producing a lower voltage from a higher
voltage, and is the basic operating mechanism of the potentiometer.
As well as being used to calculate a lower supply voltage, the voltage divider formula can also be
used in the analysis of more complex resistive circuits containing both series and parallel branches.
The voltage or potential divider formula can be used to determine the voltage drops around a closed
DC network or as part of a various circuit analysis laws such as Kirchhoff’s or Thevenin’s
theorems.

2.9.1.3 Applications of Resistors in Series


We have seen that resistors in series can be used to produce different voltages across themselves
and this type of resistor network is very useful for producing a voltage divider network. If we
replace one of the resistors in the voltage divider circuit above with a Sensor such as a thermistor,
light dependant resistor (LDR) or even a switch, we can convert an analogue quantity being sensed
into a suitable electrical signal which is capable of being measured.
For example, the following thermistor circuit has a resistance of 10KΩ at 25°C and a resistance of
100Ω at 100°C. Calculate the output voltage (Vout) for both temperatures.

Fig 2.9.3: Thermistor Circuit


At 25°C
49
𝑅2 1000
𝑉𝑜𝑢𝑡 = 𝑥 𝑉𝑖𝑛 = 𝑥 12 = 1.09𝑉
𝑅1 + 𝑅2 1000 + 1000

At 100°C
𝑅2 1000
𝑉𝑜𝑢𝑡 = 𝑥 𝑉𝑖𝑛 = 𝑥 12 = 10.9𝑉
𝑅1 + 𝑅2 100 + 1000

So by changing the fixed 1KΩ resistor, R2 in our simple circuit above to a variable resistor or
potentiometer, a particular output voltage set point can be obtained over a wider temperature range.

2.9.2 Resistors in Parallel


Unlike the previous series resistor circuit, in a parallel resistor network the circuit current can take
more than one path as there are multiple paths for the current. Then parallel circuits are classed as
current dividers.
Since there are multiple paths for the supply current to flow through, the current may not be the
same through all the branches in the parallel network. However, the voltage drop across all of the
resistors in a parallel resistive network IS the same. Then, Resistors in Parallel have a Common
Voltage across them and this is true for all parallel connected elements.
So we can define a parallel resistive circuit as one where the resistors are connected to the same
two points (or nodes) and are identified by the fact that it has more than one current path connected
to a common voltage source. Then in our parallel resistor example below the voltage across resistor
R1 equals the voltage across resistor R2 which equals the voltage across R3 and which equals the
supply voltage. Therefore, for a parallel resistor network this is given as:
𝑉𝑅1 = 𝑉𝑅2 = 𝑉𝑅3 = 𝑉𝐴𝐵 = 12𝑉
In the following resistors in parallel circuit the resistors R1, R2 and R3 are all connected together
in parallel between the two points A and B as shown in fig 2.9.4.

50
Fig 2.9.4: Parallel Resistor Circuit
In the previous series resistor network we saw that the total resistance, RT of the circuit was equal
to the sum of all the individual resistors added together. For resistors in parallel the equivalent
circuit resistance RT is calculated differently.
Here, the reciprocal ( 1/R ) value of the individual resistances are all added together instead of the
resistances themselves with the inverse of the algebraic sum giving the equivalent resistance as
shown;
Parallel Resistor Equation
𝟏 𝟏 𝟏 𝟏 𝟏
= + + …….+ 𝒆𝒕𝒄
𝑹𝑻 𝑹𝟏 𝑹𝟐 𝑹𝟑 𝑹𝒏
Then the inverse of the equivalent resistance of two or more resistors connected in parallel is the
algebraic sum of the inverses of the individual resistances.
If the two resistances or impedances in parallel are equal and of the same value, then the total or
equivalent resistance, RT is equal to half the value of one resistor. That is equal to R/2 and for three
equal resistors in parallel, R/3, etc.
Note that the equivalent resistance is always less than the smallest resistor in the parallel network
so the total resistance, RT will always decrease as additional parallel resistors are added.
Parallel resistance gives us a value known as Conductance, symbol G with the units of
conductance being the Siemens, symbol S. Conductance is the reciprocal or the inverse of
resistance, ( G = 1/R ). To convert conductance back into a resistance value we need to take the
reciprocal of the conductance giving us then the total resistance, RT of the resistors in parallel.
We now know that resistors that are connected between the same two points are said to be in
parallel. But a parallel resistive circuit can take many forms other than the obvious one given above
and here are a few examples of how resistors can be connected together in parallel.

51
Fig 2.9.5: Various Parallel Resistor Networks

The five resistive networks in fig 2.9.5 may look different to each other, but they are all arranged
as Resistors in Parallel and as such the same conditions and equations apply.

Example 1
Find the total resistance, RT of the following resistors connected in a parallel network.

The total resistance RT across the two terminals A and B is calculated as:
1 1 1 1
= + +
𝑅𝑇 𝑅1 𝑅2 𝑅3
1 1 1 1
= + + = 𝟎. 𝟎𝟏𝟏𝟕
𝑅𝑇 𝟐𝟎𝟎 𝟒𝟕𝟎 𝟐𝟐𝟎

52
1
Therefore: RT = = 85.67Ω
𝟎.𝟎𝟏𝟏𝟕

This method of reciprocal calculation can be used for calculating any number of individual
resistances connected together within a single parallel network.
If however, there are only two individual resistors in parallel then we can use a much simpler and
quicker formula to find the total or equivalent resistance value, R T and help reduce the reciprocal
maths a little.
This much quicker product-over-sum method of calculating two resistors in parallel, either having
𝑅1 𝑋 𝑅2
equal or unequal values is given as: 𝑅𝑇 = 𝑅1 +𝑅2
Example 2
Consider the following circuit which has only two resistors in a parallel combination.

Using our formula above for two resistors connected together in parallel we can calculate the total
circuit resistance, RT as:
22𝐾Ω 𝑋 47𝐾Ω
𝑅𝑇 = = 14,985Ω = 15𝑘Ω
22𝐾Ω + 47𝐾Ω
One important point to remember about resistors in parallel, is that the total circuit resistance ( RT )
of any two resistors connected together in parallel will always be LESS than the value of the
smallest resistor in that combination.
53
In our example above, the value of the combination was calculated as: RT = 15kΩ, whereas the
value of the smallest resistor is 22kΩ, much higher. In other words, the equivalent resistance of a
parallel network will always be less than the smallest individual resistor in the combination.
Also, in the case of R1 being equal to the value of R2, that is R1 = R2, the total resistance of the
network will be exactly half the value of one of the resistors, R/2.
Likewise, if three or more resistors each with the same value are connected in parallel, then the
equivalent resistance will be equal to R/n where R is the value of the resistor and n is the number
of individual resistances in the combination.
For example, six 100Ω resistors are connected together in a parallel combination. The equivalent
resistance will therefore be: RT = R/n = 100/6 = 16.7Ω. But note that this ONLY works for
equivalent resistors. That is resistors all having the same value.

2.9.2.1 Currents in a Parallel Resistor Circuit


The total current, IT entering a parallel resistive circuit is the sum of all the individual currents
flowing in all the parallel branches. But the amount of current flowing through each parallel branch
may not necessarily be the same, as the resistive value of each branch determines the amount of
current flowing within that branch.
For example, although the parallel combination has the same voltage across it, the resistances
could be different therefore the current flowing through each resistor would definitely be different
as determined by Ohms Law.
Consider the two resistors in parallel above. The current that flows through each of the resistors
( IR1 and IR2 ) connected together in parallel is not necessarily the same value as it depends upon
the resistive value of the resistor. However, we do know that the current that enters the circuit at
point A must also exit the circuit at point B.
Kirchhoff’s Current Laws states that: “the total current leaving a circuit is equal to that entering
the circuit – no current is lost“. Thus, the total current flowing in the circuit is given as:
IT = IR1 + IR2
Then by using Ohm’s Law, the current flowing through each resistor of Example 2 above can be
calculated as:
Current flowing in R1 = VS ÷ R1 = 12V ÷ 22kΩ = 0.545mA or 545μA
Current flowing in R2 = VS ÷ R2 = 12V ÷ 47kΩ = 0.255mA or 255μA
thus giving us a total current IT flowing around the circuit as:
54
IT = 0.545mA + 0.255mA = 0.8mA or 800μA
and this can also be verified directly using Ohm’s Law as:
IT = VS ÷ RT = 12 ÷ 15kΩ = 0.8mA or 800μA (the same)
The equation given for calculating the total current flowing in a parallel resistor circuit which is
the sum of all the individual currents added together is given as:
𝑰𝑻𝒐𝒕𝒂𝒍 = 𝑰𝟏 + 𝑰𝟐 + 𝑰𝟑 … . . +𝑰𝒏

Then parallel resistor networks can also be thought of as “current dividers” because the supply
current splits or divides between the various parallel branches. So a parallel resistor circuit having
N resistive networks will have N-different current paths while maintaining a common voltage
across itself. Parallel resistors can also be interchanged with each other without changing the total
resistance or the total circuit current.

Example 3
Calculate the individual branch currents and total current drawn from the power supply for the
following set of resistors connected together in a parallel combination.

As the supply voltage is common to all the resistors in a parallel circuit, we can use Ohms Law to
calculate the individual branch current as follows;
𝑉𝑆 24𝑉
𝐼1 = = = 2.4𝑎𝑚𝑝𝑠
𝑅1 10Ω

𝑉𝑆 24𝑉
𝐼2 = = = 1.2𝑎𝑚𝑝𝑠
𝑅2 20Ω

𝑉𝑆 24𝑉
𝐼3 = = = 0.8𝑎𝑚𝑝𝑠
𝑅3 30Ω

55
𝑉𝑆 24𝑉
𝐼4 = = = 0.6𝑎𝑚𝑝𝑠
𝑅4 40Ω
Then the total circuit current, IT flowing into the parallel resistor combination will be:
𝐼𝑇 = 𝐼1 + 𝐼2 + 𝐼3 + 𝐼4
𝐼𝑇 = 2.4 + 1.2 + 0.8 + 0.6
𝐼𝑇 = 5.0𝐴𝑚𝑝𝑠
This total circuit current value of 5 amperes can also be found and verified by finding the
equivalent circuit resistance, RT of the parallel branch and dividing it into the supply voltage, VS
as follows.
Equivalent circuit resistance:
𝟏 𝟏 𝟏 𝟏
= + +
𝑹𝑻 𝑹𝟏 𝑹𝟐 𝑹𝟑
𝟏 𝟏 𝟏 𝟏 𝟏
= + + +
𝑹𝑻 𝟏𝟎 𝟐𝟎 𝟑𝟎 𝟒𝟎
𝟏
= 𝟎. 𝟏 + 𝟎. 𝟎𝟓 + 𝟎. 𝟎𝟑𝟑 + 𝟎. 𝟎𝟐𝟓
𝑹𝑻
𝟏
∴ 𝑹𝑻 = = 𝟒. 𝟖Ω
𝟎. 𝟐𝟎𝟖𝟑
Then the current flowing in the circuit will be:

𝑉𝑆 24
𝐼𝑇 = = = 5𝐴𝑚𝑝𝑠
𝑅𝑇 4.8
2.10 Equivalent Resistance of Series and Parallel Circuits.
In the previous sessions, we have learnt how to connect individual resistors together to form either
a Series Resistor Network or a Parallel Resistor Network and we used Ohms Law to find the
various currents flowing in and voltages across each resistor combination.

56
This session deals with more complex resistive networks which connects various resistors together
in “BOTH” parallel and series combinations within the same circuit and how to calculate the
combined or total circuit resistance, currents and voltages for these resistive combinations.
Resistor circuits that combine series and parallel resistors networks together are generally known
as Resistor Combination or mixed resistor circuits. The method of calculating the circuit’s
equivalent resistance is the same as that for any individual series or parallel circuit and hopefully
we now know that common current flows in resistors in series and that resistors in parallel have
exactly the same voltage across them.

Example 1
In the following circuit calculate the total current ( IT ) taken from the 12v supply.

At first glance this may seem a difficult task, but if we look a little closer we can see that the two
resistors, R2 and R3 are actually both connected together in a “SERIES” combination so we can
add them together to produce an equivalent resistance the same as we did in the series resistor
calculations. The resultant resistance for this combination would therefore be:
R2 + R3 = 8Ω + 4Ω = 12Ω
So we can replace both resistors R2 and R3 above with a single resistor of resistance value 12Ω

57
So our circuit now has a single resistor RA in “PARALLEL” with the resistor R4. Using our
resistors in parallel equation we can reduce this parallel combination to a single equivalent resistor
value of R(combination) using the formula for two parallel connected resistors as follows;
1 1 1 1 1
𝑅𝑐𝑜𝑚𝑏𝑖𝑛𝑎𝑡𝑖𝑜𝑛 = + = + = = 6Ω
𝑅𝐴 𝑅4 12 12 𝑅(𝑐𝑜𝑚)
The resultant resistive circuit now looks something like this:

We can see that the two remaining resistances, R1 and R(comb) are connected together in a “SERIES”
combination and again they can be added together (resistors in series) so that the total circuit
resistance between points A and B is therefore given as:
𝑅𝑇 = 𝑅1 + 𝑅𝑐𝑜𝑚𝑏 = 6Ω + 6Ω = 12Ω

and a single resistance of just 12Ω can be used to replace the original four resistors connected
together in the original circuit.
58
Now by using Ohm´s Law, the value of the circuit current (I) is simply calculated as:
𝑉 12
𝐼= = = 1𝐴𝑚𝑝𝑒𝑟𝑒
𝑅 12
So any complicated resistive circuit consisting of several resistors can be reduced to a simple single
circuit with only one equivalent resistor by replacing all the resistors connected together in series
or in parallel using the steps above.
We can take this one step further by using Ohms Law to find the two branch currents, I1 and I2 as
shown;
𝑉𝑅𝐼 = 𝐼𝑅1 = 1 ∗ 6 = 6𝑉𝑜𝑙𝑡𝑠
𝑉𝑅𝐴 = 𝑉𝑅4 = (12 − 𝑉𝑅1 ) = 6𝑉𝑜𝑙𝑡𝑠
Thus: 𝐼1 = 6𝑉 ÷ 𝑅𝐴 = 6 ÷ 12 = 0.5𝐴 𝑜𝑟 500𝑚𝐴
𝐼2 = 6𝑉 ÷ 𝑅4 = 6 ÷ 12 = 0.5𝐴 𝑜𝑟 500𝑚𝐴
Since the resistive values of the two branches are the same at 12Ω, the two branch currents of I1
and I2 are also equal at 0.5A (or 500mA) each. This therefore gives a total supply current,
IT of: 0.5 + 0.5 = 1.0 amperes (as calculated above.)
It is sometimes easier with complex resistor combinations and resistive networks to sketch or
redraw the new circuit after these changes have been made, as this helps as a visual aid. Then
continue to replace any series or parallel combinations until one equivalent resistance, REQ is
found.
Example 2
Find the equivalent resistance, REQ for the following resistor combination circuit.

Again, at first glance this resistor ladder network may seem a complicated task, but as before it is
just a combination of series and parallel resistors connected together. Starting from the right

59
hand side and using the simplified equation for two parallel resistors, we can find the equivalent
resistance of the R8 to R10 combination and call it RA.

RA is in series with R7 therefore the total resistance will be RA + R7 = 4 + 8 = 12Ω as shown.

This resistive value of 12Ω is now in parallel with R6 and can be calculated as RB.

RB is in series with R5 therefore the total resistance will be RB + R5 = 4 + 4 = 8Ω as shown.

60
This resistive value of 8Ω is now in parallel with R4 and can be calculated as RC as shown.

RC is in series with R3 therefore the total resistance will be RC + R3 = 8Ω as shown.

This resistive value of 8Ω is now in parallel with R2 from which we can calculated RD as:

RD is in series with R1 therefore the total resistance will be RD + R1 = 4 + 6 = 10Ω as shown.

61
Then the complex combinational resistive network above comprising of ten individual resistors
connected together in series and parallel combinations can be replaced with just one single
equivalent resistance (REQ) of value 10Ω.
When solving any combinational resistor circuit that is made up of resistors in series and parallel
branches, the first step we need to take is to identify the simple series and parallel resistor branches
and replace them with equivalent resistors.
This step will allow us to reduce the complexity of the circuit and help us transform a complex
combinational resistive circuit into a single equivalent resistance remembering that series circuits
are voltage dividers and parallel circuits are current dividers.
However, calculations of more complex T-pad Attenuator and resistive bridge networks which
cannot be reduced to a simple parallel or series circuit using equivalent resistances require a
different approach. These more complex circuits need to be solved using Kirchhoff’s Current Law,
and Kirchhoff’s Voltage Law.

2.11 Potential Difference


Unlike current which flows around a closed electrical circuit in the form of electrical charge,
potential difference does not move or flow it is applied. The unit of potential difference generated
between two points is called the Volt and is generally defined as being the potential difference
dropped across a fixed resistance of one ohm with a current of one ampere flowing through it.
In other words, 1 Volt equals 1 Ampere times 1 Ohm, or commonly V = I*R.
Ohm’s Law states that for a linear circuit the current flowing through it is proportional to the
potential difference across it so the greater the potential difference across any two points the bigger
will be the current flowing through it.
For example, if the voltage at one side of a 10Ω resistor measures 8V and at the other side of the
resistor it measures 5V, then the potential difference across the resistor would be 3V ( 8 – 5 )
causing a current of 0.3A to flow.
If however, the voltage on one side was increased from 8V to say 40V, the potential difference
across the resistor would now be 40V – 5V = 35V causing a current of 3.5A to flow. The voltage
at any point in a circuit is always measured with respect to a common point, generally 0V.
For electrical circuits, the earth or ground potential is usually taken to be at zero volts ( 0V ) and
everything is referenced to that common point in a circuit. This is similar in theory to measuring

62
height. We measure the height of hills in a similar way by saying that the sea level is at zero feet
and then compare other points of the hill or mountain to that level.
In a very similar way we can call the common point in a circuit zero volts and give it the name of
ground, zero volts or earth, then all other voltage points in the circuit are compared or referenced
to that ground point. The use of a common ground or reference point in electrical schematic
drawings allows the circuit to be drawn more simply as it is understood that all connections to this
point have the same potential. For example:
As the units of measure for Potential Difference are volts, potential difference is mainly called
voltage. Individual voltages connected in series can be added together to give us a “total voltage”
sum of the circuit as seen in the resistors in series. Voltages across components that are connected
in parallel will always be of the same value as seen in the resistors in parallel, for example.
For voltages connected in series:
𝑉𝑇 = 𝑉1 + 𝑉2 + 𝑉3 … 𝑒𝑡𝑐
For parallel connected voltages:
𝑉𝑇 = 𝑉1 = 𝑉2 = 𝑉3 … 𝑒𝑡𝑐

Example 1
Calculate the current flowing through a 100Ω resistor that has one of its terminals connected to 50
volts and the other terminal connected to 30 volts.
Solution:
Voltage at terminal A is equal to 50v and the voltage at terminal B is equal to 30v. Therefore, the
voltage across the resistor is given as:
VA = 50v, VB = 30v,
Therefore, VA – VB = 50 – 30 = 20v
Using ohms law, I = VAB ÷ R = 20V ÷ 100Ω = 200mA

2.11.1 Voltage Divider Network


We know from the previous sessions that by connecting together resistors in series across a
potential difference we can produce a voltage divider circuit which will give the ratios of voltages
across each resistor with respect to the supply voltage across the total combination.
This produces what is generally called a Voltage Divider Network and one which only applies to
resistors connected together in series, because as we saw in the Resistors in Parallel, resistors

63
connected together in parallel produce what is called a current divider network. Consider the series
circuit below.

Fig 2.6: Voltage Division


The circuit shows the principle of a voltage divider circuit where the output voltage drops across
each resistor within the series chain, with resistors R1, R2, R3 and R4 being referenced to some
common reference point (usually zero volts).
So for any number of resistors connected together in series, dividing the supply voltage VS by the
total resistance RT, will give the current flowing through the series branch as: I = VS/RT, (Ohm’s
Law). Then the individual voltage drops across each resistor can be simply calculated as: V = I*R
where R represents the resistance value.
The voltage at each point, P1, P2, P3 etc. increases according to the sum of the voltages at each
point up to the supply voltage, Vs and we can also calculate the individual voltage drops at any
point without firstly calculating the circuit current by using the following formula;
Voltage Divider Formula
𝑹(𝒙)
𝑽𝒙 = 𝑽
𝑹𝑻 𝒔
Where, V(x) is the voltage to be found, R(x) is the resistance producing the voltage, RT is the total
series resistance and VS is the supply voltage.
Example 2
In the circuit above, four resistors of values, R1 = 10Ω, R2 = 20Ω, R3 = 30Ω and R4 = 40Ω are
connected across a 100 volts DC supply. Using the formula above, calculated the voltage drops at
points P1, P2, P3 and P4 and also the individual voltage drops across each resistor within the series
chain.
64
1. The voltages at the various points are calculated as:

2. The individual voltage drops across each resistor are calculated as:

Then by using this equation we can say that the voltage dropped across any resistor in a series
circuit is proportional to the magnitude of the resistor and the total voltage dropped across all the
resistors must equal the voltage source as defined by Kirchhoff’s Voltage Law. So by using the
Voltage Divider Equation, for any number of series resistors the voltage drop across any
individual resistor can be found.
Thus far we have seen that voltage is applied to a resistor or circuit and that current flows through
and around a circuit. But there is a third variable we can also apply to resistors and resistor

65
networks. Power is a product of voltage and current and the basic unit of measurement of power
is the watt.
In the next tutorial about Resistors, we will examine the power dissipated (consumed) by resistance
in the form of heat and that the total power dissipated by a resistive circuit, whether it is series,
parallel, or a combination of the two, we simply add the powers dissipated by each resistor.

2.11.2 Resistor Power Rating


When an electrical current passes through a resistor due to the presence of a voltage across it,
electrical energy is lost by the resistor in the form of heat and the greater this current flow the
hotter the resistor will get. This is known as the Resistor Power Rating.
Resistors are rated by the value of their resistance and the electrical power given in watts, (W) that
they can safely dissipate based mainly upon their size. Every resistor has a maximum power rating
which is determined by its physical size as generally, the greater its surface area the more power
it can dissipate safely into the ambient air or into a heatsink.
A resistor can be used at any combination of voltage (within reason) and current so long as its
“Dissipating Power Rating” is not exceeded with the resistor power rating indicating how much
power the resistor can convert into heat or absorb without any damage to itself. The Resistor
Power Rating is sometimes called the Resistors Wattage Rating and is defined as the amount of
heat that a resistive element can dissipate for an indefinite period of time without degrading its
performance.
The power rating of resistors can vary a lot from less than one tenth of a watt to many hundreds of
watts depending upon its size, construction and ambient operating temperature. Most resistors have
their maximum resistive power rating given for an ambient temperature of +70oC or below.
Electrical power is the rate in time at which energy is used or consumed (converted into heat). The
standard unit of electrical power is the Watt, symbol W and a resistors power rating is also given
in Watts. As with other electrical quantities, prefixes are attached to the word “Watt” when
expressing very large or very small amounts of resistor power. Some of the more common of these
are:

66
Electrical Power Units
Unit Symbol Value Abbreviation

Milliwatt mW 1/1,000th watt 10-3 W

Kilowatt kW 1,000 watts 103 W

Megawatt MW 1,000,000 watts 106 W

From Ohm’s Law a current flows through a resistance, a voltage is dropped across it producing a
product which relates to power. In other words, if a resistance is subjected to a voltage, or if it
conducts a current, then it will always consume electrical power and we can superimpose these
three quantities of power, voltage and current into a triangle called a Power Triangle with the
power, which would be dissipated as heat in the resistor at the top, with the current consumed and
the voltage across it at the bottom as shown.

Fig 2.7: The Resistor Power Triangle


The above power triangle is great for calculating the power dissipated in a resistor if we know the
values of the voltage across it and the current flowing through it. But we can also calculate the
power dissipated by a resistance by using Ohm’s Law.
Ohms law allows us to calculate the power dissipation given the resistance value of the resistor.
By using Ohms Law it is possible to obtain two alternative variations of the above expression for

67
the resistor power if we know the values of only two, the voltage, the current or the resistance as
follows:
[P = V x I] Power = Volts x Amps
[P = I2 x R] Power = Current2 x Ohms
[P = V2 ÷ R] Power = Volts2 ÷ Ohms
The electrical power dissipation of any resistor in a DC circuit can be calculated using one of the
following three standard formulas:
𝑽𝟐
𝑷𝒐𝒘𝒆𝒓 (𝑷) = 𝑽 × 𝑰 = 𝑰𝟐 𝑹 =
𝑹
Where: V is the voltage across the resistor in Volts
I is in current flowing through the resistor in Amperes
R is the resistance of the resistor in Ohm’s (Ω)
As the dissipated resistor power rating is linked to their physical size, a 1/4 (0.250)W resistor is
physically smaller than a 1W resistor, and resistors that are of the same ohmic value are also
available in different power or wattage ratings. Carbon resistors, for example, are commonly made
in wattage ratings of 1/8 (0.125)W, 1/4 (0.250)W, 1/2 (0.5)W, 1W, and 2 Watts.
Generally speaking the larger their physical size the higher its wattage rating. However, it is always
better to select a particular size resistor that is capable of dissipating two or more times the
calculated power. When resistors with higher wattage ratings are required, wirewound resistors
are generally used to dissipate the excessive heat.
Type Power Rating Stability

Metal Film Very low at less than 3 Watts High 1%

Carbon Low at less than 5 Watts Low 20%

Wirewound High up to 500 Watts High 1%

Example 1
What is the maximum power rating in watts of a fixed resistor which has a voltage of 12 volts
across its terminals and a current of 50 milliamperes flowing through it.
Given that we know the values of the voltage and current above, we can substitute these values
into the following equation: P = V*I.

68
Example 2
Calculate the maximum safe current that can pass through a 1.8KΩ resistor rated at 0.5 Watts.
Again, as we know the resistors power rating and its resistance, we can now substitute these values
into the standard power equation of:
𝑃 = 𝐼2 𝑅

𝑃
𝐼=√
𝑅

0.5
=√ = 0.016A or 16mA
1800

All resistors have a Maximum Dissipated Power Rating, which is the maximum amount of
power it can safely dissipate without damage to itself. Resistors which exceed their maximum
power rating tend to go up in smoke, usually quite quickly, and damage the circuit they are
connected to. If a resistor is to be used near to its maximum power rating then some form of heat
sink or cooling is required.
Resistor power rating is an important parameter to consider when choosing a resistor for a
particular application. The job of a resistor is to resist current flow through a circuit and it does
this by dissipating the unwanted power as heat. Selecting a small wattage value resistor when high
power dissipation is expected will cause the resistor to overheat, destroying both the resistor and
the circuit.
Thus far we have considered resistors connected to a steady DC supply, but in the next tutorial
about Resistors, we will look at the behaviour of resistors that are connected to a sinusoidal AC
supply, and show that the voltage, current and therefore the power consumed by a resistor used in
an AC circuit are all in-phase with each other.

69
MODULE THREE

3.1 Electric Circuits and Network Theorems


There are certain theorems, which when applied to the solutions of electric networks, wither
simplify the network itself or render their analytical solution very easy. These theorems can also
be applied to an a.c. system, with the only difference that impedances replace the ohmic resistance
of d.c. system. Different electric circuits (according to their properties) are defined below:
3.1 1. Circuit. A circuit is a closed conducting path through which an electric current either flows
or is intended flow.
3.1 2. Parameters. The various elements of an electric circuit are called its parameters like
resistance, inductance and capacitance. These parameters may be lumped or distributed.
3.1 3. Liner Circuit. A linear circuit is one whose parameters are constant i.e.they do not change
with voltage or current.
3.1 4. Non-linear Circuit. It is that circuit whose parameters change with voltage or current.
3.1 5. Bilateral Circuit. A bilateral circuit is one whose properties or characteristics are the same
in either direction. The usual transmission line is bilateral, because it can be made to perform its
function equally well in either direction.
3.1 6. Unilateral Circuit. It is that circuit whose properties or characteristics change with the
direction of its operation. A diode rectifier is a unilateral circuit, because it cannot perform
rectification in both directions.
3.1 7. Electric Network. A combination of various electric elements, connected in any manner
whatsoever, is called an electric network.
3.1 8. Passive Network is one which contains no source of e.m.f. in it.
3.1 9. Active Network is one which contains one or more than one source of e.m.f.
3.1 10. Node is a junction in a circuit where two or more circuit elements are connected together.
3.1 11. Branch is that part of a network which lies between two junctions.
3.1 12. Loop. It is a close path in a circuit in which no element or node is encountered more than
once
3.1 13. Mesh. It is a loop that contains no other loop within it. For example, the circuit of Fig.3.1
(a) has even branches, six nodes, three loops and two meshes whereas the circuit of Fig. 3.1 (b)
has four branches, two nodes, six loops and three meshes.

70
Figure 3.1
The laws which determine the currents and voltage drops in d.c. networks are:
(a) Ohm’s law
(b) the laws for resistors in series and in parallel and
(c) Kirchhoff’s laws
In addition, there are a number of circuit theorems which have been developed for solving
problems in electrical networks. These include: superposition theorem, Thévenin’s theorem
Norton’s theorem and the maximum power transfer theorem.
3.2 Kirchhoff’s laws
From our consideration of series and of parallel connections of resistors, we have observed certain
conditions appertaining to each form of connection. For instance, in a series circuit, the sum of the
voltages across each of the components is equal to the applied voltage; again the sum of the
currents in the branches of a parallel network is equal to the supply current. Gustav Kirchhoff, a
German physicist, observed that these were particular instances of two general conditions
fundamental to the analysis of any electrical network.
3.2.1 Kirchhoff’s Current Law
States that at any junction in an electric circuit the total current flowing towards that junction is
equal to the total current flowing away from the junction,
i.e. ∑ 𝐼 = 0
Thus,

71
Figure 3.2

3.2.2 Kirchhoff’s Voltage Law.


States that in any closed loop in a network, the algebraic sum of the voltage drops (i.e. products
of current and resistance) taken around the loop is qual to the resultant e.m.f. acting in that loop.
Thus

Figure 3.3

(Note that if current flows away from the positive terminal of a source, that source is considered
by convention to be positive. Thus moving anticlockwise around the loop of Fig. E1 is positive
and E2 is negative).
Example 1.
(a) Find the unknown currents marked in Fig. (a).
(b) Determine the value of e.m.f. E in Fig.(b).

72
73
Example 2
Determine, using Kirchhoff’s laws, each branch current for the network shown in figure below

Figure 3.4

74
Solution

Example 3
Determine, using Kirchhoff’s laws, each branch current for the network shown in figure

Figure 3.5

75
Solution
Currents, and their directions are shown labelled in Fig. 13.8 following Kirchhoff’s current law. It
is usual, although not essential, to follow conventional current flow with current flowing from the
positive terminal of the source

Figure 3.6

The network is divided into two loops as shown in Fig. 13.8. Applying Kirchhoff’s voltage law
gives: For loop 1:

76
Exercise

3.3 The superposition theorem


The superposition theorem states:
In any network made up of linear resistances and containing more than one source of e.m.f., the
resultant current flowing in any branch is the algebraic sum of the currents that would flow in that
branch if each source was considered separately, all other sources being replaced at that time by
their respective internal resistances.
Example 1.
Figure 13.16 shows a circuit containing two sources of e.m.f., each with their internal resistance.
Determine the current in each branch of the network by using the superposition theorem.

77
Figure 3.7

1. Redraw the original circuit with source E2 removed, being replaced by r2 only, as shown in Fig.
3.7(a)

Figure 3.

2. Label the currents in each branch and their directions as shown in Fig. 13.17(a) and determine
their values. (Note that the choice of current directions depends on the battery polarity, which, by
convention is taken as flowing from the positive battery terminal as shown)

78
79
Figure 1: An Atom

Figure 1: An Atom

80
Figure 1: An Atom

81
MODULE FOUR

4.1 Temperature coefficient of resistance


In general, as the temperature of a material increases, most conductors increase in resistance,
insulators decrease in resistance, whilst the resistance of some special alloys remain almost
constant. The resistance of all pure metals increases with increase of temperature, whereas the
resistance of carbon, electrolytes and insulating materials decreases with increase of temperature.
Certain alloys, such as manganin, show practically no change of resistance for a considerable
variation of temperature.
The temperature coefficient of resistance of a material is the increase in the resistance of a 1𝛺
resistor of that material when it is subjected to a rise of temperature of 1◦C. The symbol used for
the temperature coefficient of resistance is α (Greek alpha). Thus, if some copper wire of resistance
1𝛺 is heated through 1◦C and its resistance is then measured as 1.0043𝛺 then α=0.0043 𝛺 / 𝛺◦C
for copper. The units are usually expressed only as ‘per ◦C’, i.e. α=0.0043/◦C for copper. If the 1
𝛺 resistor of copper is heated through 100◦C then the resistance at 100◦C would be
1+100×0.0043=1.43 𝛺. Some typical values of temperature coefficient of resistance measured at
0◦C are given below:
Copper 0.0043/◦C
Nickel 0.0062/◦C
Constantan 0
Aluminium 0.0038/◦C
Carbon −0.00048/◦C
Eureka 0.00001/◦C
(Note that the negative sign for carbon indicates that its resistance falls with increase of
temperature.)
Table: Typical temperature coefficients of resistance referred to 0 °C

82
4.2 Variation of resistance of copper with temperature

Figure 4.1: Variation of resistance of copper with temperature


The variation of resistance of copper for a range over which copper conductors are usually operated
is represented by the graph in Fig. 3.40. If this graph is extended backwards, the point of
intersection with the horizontal axis is found to be –234.5 °C.
If the resistance of a material at 0◦C is known the resistance at any other temperature can be
determined from:
Rθ =R0(1+α0θ)
where
R0 = resistance at 0◦C

83
Rθ = resistance at temperature θ ◦C
α0 = temperature coefficient of resistance at 0◦C

Worked examples
Example 1
A coil of copper wire has a resistance of 100 𝛺 when its temperature is 0◦C. Determine its
resistance at 70◦C if the temperature coefficient of resistance of copper at 0◦C is 0.0043/ ◦C.
Resistance Rθ =R0 (1+α0θ). Hence resistance at
100◦C,
R100 = 100[1 + (0.0043)(70)]
= 100[1 + 0.301]
= 100(1.301) = 130.1 𝛺
Example 2
An aluminium cable has a resistance of 27 𝛺 at a temperature of 35◦C. Determine its resistance at
0◦C. Take the temperature coefficient of resistance at 0◦C to be 0.0038/ ◦C.
Resistance at θ ◦C,
Rθ =R0 (1+α0θ). Hence resistance
at 0◦C,
R0 = Rθ (1 + α0θ) = 27
[1 + (0.0038)(35)] = 27

84
1 + 0.133 = 27
1.133
= 23.83 𝛺
Example 3
A carbon resistor has a resistance of 1 k 𝛺 at 0◦C. Determine its resistance at 80◦C. Assume that
the temperature coefficient of resistance for carbon at 0◦C is −0.0005/ ◦C. Resistance at
temperature θ ◦C,
Rθ = R0 (1 + α0θ)
i.e.
Rθ = 1000[1 + (−0.0005)(80)]
= 1000[1 − 0.040]
= 1000(0.96)
= 960 𝛺
If the resistance of a material at room temperature (approximately 20◦C), R20, and the temperature
coefficient of resistance at 20◦C, α20, are known then the resistance Rθ at temperature θ ◦C is
given by:
Rθ =R20 [1+α20(θ −20)]
Example 4
A coil of copper wire has a resistance of 10_ at 20◦C. If the temperature coefficient of resistance
of copper at 20◦C is 0.004/ ◦C determine the resistance of the coil when the temperature rises to
100◦C.
Resistance at θ ◦C,
Rθ = R20 [1 + α20(θ − 20)]
Hence resistance at 100◦C,
R100 = 10[1 + (0.004)(100 − 20)]
= 10[1 + (0.004)(80)]
= 10[1 + 0.32]
= 10(1.32)
= 13.2 𝛺

85
Example 5
The resistance of a coil of aluminium wire at 18◦C is 200 𝛺. The temperature of the wire is
increased and the resistance rises to 240 𝛺. If the temperature coefficient of resistance of
aluminium is 0.0039/ ◦C at 18◦C determine the temperature to which the coil has risen.
Let the temperature rise to θ ◦C. Resistance at θ ◦C,
Rθ = R18 [1 + α18(θ − 18)]
i.e.
240 = 200[1 + (0.0039)(θ − 18)]
240 = 200 + (200)(0.0039)(θ − 18)
240 − 200 = 0.78(θ − 18)
40 = 0.78(θ − 18)
40 = 0.78(θ – 18)
51.28 = θ − 18, from which,
θ = 51.28 + 18
= 69.28◦C
Example 6
When a potential difference of 10 V is applied to a coil of copper wire of mean temperature 20 °C,
a current of 1.0 A flows in the coil. After some time the current falls to 0.95 A yet the supply
voltage remains unaltered. Determine the mean temperature of the coil given that the temperature
coefficient of resistance of copper is 4.28 × 10−3/°C referred to 0 °C.

86
Solution

When an electrical device is loaded (e.g. when supplying electrical power in the case of a
generator, mechanical power in the case of a motor or acting as an amplifier in the case of a
transistor), the temperature rise of the device is largely due to the I2R loss in the conductors; and
the greater the load, the greater is the loss and therefore the higher the temperature rise. The full
load or rated output of the device is the maximum output obtainable under specified conditions,
e.g. for a specified temperature rise after the device has been loaded continuously for a period of
minutes or hours. The temperature of a coil can be measured by the following means:
1. A thermometer.
2. The increase of resistance of the coil.
3. Thermo-junctions embedded in the coil.
The third method enables the distribution of temperature throughout the coil to be determined, but
is only possible if the thermo-junctions are inserted in the coil when the latter is being wound.
Since the heat generated at the centre of the coil has to flow outwards, the temperature at the centre
may be considerably higher than that at the surface. The temperature of an electronic device,
especially one incorporating a semiconductor junction, is of paramount importance, since even a
small rise of temperature above the maximum permissible level rapidly leads to a catastrophic
breakdown.

87
Exercise
1. Explain what is meant by the temperature coefficient of resistance of a material. A copper rod,
0.4 m long and 4.0 mm in diameter, has a resistance of 550 μΩ at 20 °C. Calculate the resistivity
of copper at that temperature. If the rod is drawn out into a wire having a uniform diameter of 0.8
mm, calculate the resistance of the wire when its temperature is 60 °C. Assume the resistivity to
be unchanged and the temperature coefficient of resistance of copper to be 0.004 26/°C. [0.0173

2. A coil of insulated copper wire has a resistance of 150 Ω at 20 °C. When the coil is connected
across a 230 V supply, the current after several hours is 1.25 A. Calculate the average temperature
throughout the coil, assuming the temperature coefficient of resistance of copper at 20 °C to be
0.0039/°C.[ 78.1°C]
3. A copper cable has a resistance of 30_ at a temperature of 50◦C. Determine its resistance at 0◦C.
Take the temperature coefficient of resistance of copper at 0◦C as 0.0043/◦C [24.69_]
3. The temperature coefficient of resistance for carbon at 0◦C is −0.00048/ ◦C. What is the
significance of the minus sign? A carbon resistor has a resistance of 500_at 0◦C. Determine its
resistance at 50◦C. [488_]
4. A coil of copper wire has a resistance of 20_ at 18◦C. If the temperature coefficient of resistance
of copper at 18◦Cis0.004/ ◦C, determine the resistance of the coil when the temperature rises to
98◦C [26.4_]
5. The resistance of a coil of nickel wire at 20◦C is 100_. The temperature of the wire is increased
and the resistance rises to 130_. If the temperature coefficient of resistance of nickel is 0.006/ ◦C
at 20◦ C, determine the temperature to which the coil has risen. [70◦C]
6. Some aluminium wire has a resistance of 50_ at 20◦C. The wire is heated to a temperature of
100◦C. Determine the resistance of the wire at 100◦C, assuming that the temperature coefficient of
resistance at 0◦C is 0.004/ ◦C. [64.8_]

88
MODULE FIVE
5.1 Capacitance and Capacitors

A capacitor is a device that can store electric charge for short periods of time. Like resistors,
capacitors can be connected in series and in parallel and therefore we can analyse them after the
fashion which we have developed in previous chapters. We know that a resistor makes the passage
of electric charge difficult, hence the production of heat, but otherwise we do not bother too much
about what happens in the resistor. However, the effect of storing charge in a capacitor has much
more significance not only within the capacitor but also in the space surrounding it. The effect in
such space is termed the electric field and this requires that we investigate it in some detail. If we
wished to fill a container with water, we know that it takes time to pour in that water. In much the
same way, it takes time to pour charge into a capacitor and again this speed of action is something
with which we need to become familiar since it has a great deal of influence on the application of
capacitors. Capacitors are widely used in all branches of electrical engineering and the effect of
capacitance is to be found wherever there is an electric circuit. We shall find that capacitors are
one of the three main components in any electrical system.

A capacitor is an electrical device that is used to store electrical energy. Next to the resistor, the
capacitor is the most commonly encountered component in electrical circuits. Capacitors are used
extensively in electrical and electronic circuits. For example, capacitors are used to smooth
rectified a.c. outputs, they are used in telecommunication equipment – such as radio receivers –
for tuning to the required frequency, they are used in time delay circuits, in electrical filters, in
oscillator circuits, and in magnetic resonance imaging (MRI) in medical body scanners, to name
but a few practical applications.

89
5.2 Electrostatic field

Figure 5.1
Figure 5.1 represents two parallel metal plates, A and B, charged to different potentials. If an
electron that has a negative charge is placed between the plates, a force will act on the electron
tending to push it away from the negative plate B towards the positive plate, A. Similarly, a positive
charge would be acted on by a force tending to move it toward the negative plate.
Any region such as that shown between the plates in Fig. 5.1, in which an electric charge
experiences a force, is called an electrostatic field. The direction of the field is defined as that of
the force acting on a positive charge placed in the field. In Fig. 5.1, the direction of the force
is from the positive plate to the negative plate. Such a field may be represented in magnitude and
direction by lines of electric force drawn between the charged surfaces. The closeness of the lines
is an indication of the field strength. Whenever a p.d. is established between two points, an electric
field will always exist.
Figure 5.2(a) shows a typical field pattern for an isolated point charge, and Fig. 5.2(b) shows the
field pattern for adjacent charges of opposite polarity. Electric lines of force (often called electric
flux lines) are continuous and start and finish on point charges; also, the lines cannot cross each
other. When a charged body is placed close to an uncharged body, an induced charge of opposite
sign appears on the surface of the uncharged body. This is because lines of force from the charged
body terminate on its surface. The concept of field lines or lines of force is used to illustrate the
properties of an electric field. However, it should be remembered that they are only aids to the
imagination. The force of attraction or repulsion between two electrically charged bodies is
proportional to the magnitude of their charges and inversely proportional to the square of the
distance separating them, i.e.

90
Figure 5.2

91
Solution

92
93
The space surrounding a charge can be investigated using a small charged body. This investigation
is similar to that applied to the magnetic field surrounding a current-carrying conductor. However,
in this case the charged body is either attracted or repelled by the charge under investigation. The
space in which this effect can be observed is termed the electric field of the charge and the force
on the charged body is the electric force. The lines of force can be traced out and they appear to
have certain properties:
1. In an electric field, each line of force emanates from or terminates in a charge. The conventional
direction is from the positive charge to the negative charge.
2. The direction of the line is that of the force experienced by a positive charge placed at a point
in the field, assuming that the search charge has no effect on the field which it is being used to
investigate.
3. The lines of force never intersect since the resultant force at any point in the field can have only
one direction.
The force of attraction or of repulsion acts directly between two adjacent charges. All points on
the surface of a conductor may be assumed to be at the same potential, i.e. equipotential, and the
lines of force radiate out from equipotential surfaces at right angles. The simplest case is that of
the isolated spherical charge. However, most electric fields exist between two conductors. The two
most important arrangements are those involving parallel plates (as in a simple capacitor) and
concentric cylinders (as in a television aerial cable). The resulting fields are shown in Fig. 5.3.

Figure 5. 3 Electric fields between oppositely charged surfaces. (a) Parallel plates; (b) concentric
cylinders (cable)
5.3 Electric field strength
Figure 5.3 shows two parallel conducting plates separated from each other by air. They are
connected to opposite terminals of a battery of voltage V volts. There

94
Figure 5.4
is therefore an electric field in the space between the plates. If the plates are close together, the
electric lines of force will be straight and parallel and equally spaced, except near the edge where
fringing will occur Over the area in which there is negligible fringing,

Where d is the distance between the plates. Electric field strength is also called potential gradient.
5.4 Capacitance
Static electric fields arise from electric charges, electric field lines beginning and ending on electric
charges. Thus the presence of the field indicates the presence of equal positive and negative electric
charges on the two plates of Fig. 5.4 Let the charge be +Q coulombs on one plate and −Q coulombs
on the other. The property of this pair of plates which determines how much charge corresponds
to a given p.d. between the plates is called their capacitance:

The unit of capacitance is the farad F (or more usually μF=10−6 F or pF =10−12 F), which is
defined as the capacitance when a p.d. of one volt appears across the plates when charged with one
coulomb.

95
5.5 Capacitors
Every system of electrical conductors possesses capacitance. For example, there is capacitance
between the conductors of overhead transmission lines and also between the wires of a telephone
cable. In these examples the capacitance is undesirable but has to be accepted, minimised or
compensated for. There are other situations where capacitance is a desirable property. Devices
specially constructed to possess capacitance are called capacitors (or condensers, as they used to
be called). In its simplest form a capacitor consists of two plates which are separated by an
insulating material known as a dielectric. A capacitor has the ability to store a quantity of static
electricity. The symbols for a fixed capacitor and a variable capacitor used in electrical circuit
diagrams are shown in Fig. 5.5

Figure 5.5

The charge Q stored in a capacitor is given by:

5.6 COMBINATIONS OF CAPACITORS


Two or more capacitors can be combined in circuits in several ways, but most reduce to two simple
configurations, called parallel and series. The idea, then, is to find the single equivalent
capacitance due to a combination of several different capacitors that are in parallel or in series with
each other. Capacitors are manufactured with a number of different standard capacitances, and by
combining them in different ways, any desired value of the capacitance can be obtained.
5.6.1 Capacitors in Parallel
Two capacitors connected as shown in Active Figure 5.6a are said to be in parallel. The left plate
of each capacitor is connected to the positive terminal of the battery by a conducting wire, so the
left plates are at the same potential. In the same way, the right plates, both connected to the negative
terminal of the battery, are also at the same potential. This means that capacitors in parallel both

96
have the same potential difference _V across them. Capacitors in parallel are illustrated in
Figure 5.6b. When the capacitors are first connected in the circuit, electrons are transferred from
the left plates through the battery to the right plates, leaving the left plates positively charged and
the right plates negatively charged. The energy source for this transfer of charge is the internal
chemical energy stored in the battery, which is converted to electrical energy. The flow of charge
stops when the voltage across the capacitors equals the voltage of the battery, at which time the
capacitors have their maximum charges. If the maximum charges on the two capacitors are Q1 and
Q 2, respectively, then the total charge, Q, stored by the two capacitors is

We can replace these two capacitors with one equivalent capacitor having a capacitance of Ceq.
This equivalent capacitor must have exactly the same external effect on the circuit as the original
two, so it must store Q units of charge and have the same potential difference across it. The
respective charges on each capacitor are

97
Figure 5.6

(a) A parallel connection of two FIGURE 1 capacitors. (b) The circuit diagram for the parallel
combination. (c) The potential differences across the capacitors are the same, and the equivalent
capacitance is Ceq = C1 + C2.
We see that the equivalent capacitance of a parallel combination of capacitors is greater than any
of the individual capacitances.
5.6.2 Capacitors in Series
For a series combination of capacitors, the magnitude of the charge must be the same on all the
plates. To understand this principle, consider the charge transfer process in some detail. When a
battery is connected to the circuit, electrons with total charge _Q are transferred from the left plate
of C1 to the right plate of C2 through the battery, leaving the left plate of C1 with a charge of _Q.
As a consequence, the magnitudes of the charges on the left plate of C1 and the right plate of C2
must be the same. Now consider the right plate of C1 and the left plate of C2, in the middle. These
98
plates are not connected to the battery (because of the gap across the plates) and, taken together,
are electrically neutral. The charge of _Q on the left plate of C1, however, attracts negative charges
to the right plate of C1. These charges will continue to accumulate until the left and right plates of
C1, taken together, become electrically neutral, which means the charge on the right plate of C1 is
_Q. This negative charge could only have come from the left plate of C2, so C2 has a charge of
_Q.

Figure 5.7

A series combination of two capacitors. The charges on the capacitors are the same, and the
equivalent capacitance can be calculated from the reciprocal relationship
1 1 1
= +
𝐶𝑒𝑞 𝐶1 𝐶2
Therefore, regardless of how many capacitors are in series or what their capacitances are, all of
the right plates gain charges of _Q and all the left plates have charges of _Q. (This is a
consequence of the conservation of charge.) After an equivalent capacitor for a series of capacitors
is fully charged, the equivalent capacitor must end up with a charge of _Q on its right plate
and a charge of _Q on its left plate. Applying the definition of capacitance to the circuit
in Active Figure 5.7b, we have

99
Where ∆V is the potential difference between the terminals of the battery and Ceq is the equivalent
capacitance. Because Q =C∆V can be applied to each capacitor, the potential differences across
them are given by

The potential difference across any number of capacitors (or other circuit elements) in series equals
the sum of the potential differences across the individual capacitors.

Worked examples
Problem 1. (a) Determine the p.d. across a 4 μF capacitor when charged with 5mC (b) Find the
charge on a 50 pF capacitor when the voltage applied to it is 2 kV.

100
Problem 2. A direct current of 4A flows into a previously uncharged 20 μF capacitor for 3 ms.
Determine the p.d. between the plates.
Solution:

Exercise
1. A 20 μF capacitor is charged at a constant current of 5 μA for 10 min. Calculate the final p.d.
across the capacitor and the corresponding charge in coulombs.

101
2. Three capacitors have capacitances of 10 μF, 15 μF and 20 μF respectively. Calculate the total
capacitance when they are connected (a) in parallel, (b) in series.
3. A 9 μF capacitor is connected in series with two capacitors, 4 μF and 2 μF respectively, which
are connected in parallel. Determine the capacitance ofthe combination. If a p.d. of 20 V is
maintained across the combination, determine the charge on the 9 μF capacitor and the energy
stored in the 4 μF capacitor.
4. Two capacitors, having capacitances of 10 μF and 15 μF respectively, are connected in series
across a 200 V d.c. supply. Calculate: (a) the charge on each capacitor; (b) the p.d. across each
capacitor. Also find the capacitance of a single capacitor that would be equivalent to these two
capacitors in series.
5. Three capacitors of 2, 3 and 6 μF respectively are connected in series across a 500 V d.c. supply.
Calculate: (a) the charge on each capacitor; (b) the p.d. across each capacitor; and (c) the energy
stored in the 6 μF capacitor.
6. A certain capacitor has a capacitance of 3 μF. A capacitance of 2.5 μF is required by combining
this capacitance with another. Calculate the capacitance of the second capacitor and state how it
must be connected to the first.
7. A capacitor A is connected in series with two capacitors B and C connected in parallel. If the
capacitances of A, B and C are 4, 3 and 6 μF respectively, calculate the equivalent capacitance of
the combination. If a p.d. of 20 V is maintained across the whole circuit, calculate
the charge on the 3 μF capacitor.
8. Three capacitors, A, B and C, are connected in series across a 200 V d.c. supply. The p.d.s across
the capacitors are 40, 70 and 90 V respectively. If the capacitance of A is 8 μF, what are the
capacitances of B and C?
9. Two capacitors, A and B, are connected in series across a 200 V d.c. supply. The p.d. across A
is 120 V. This p.d. is increased to 140 V when a 3 μF capacitor is connected in parallel with B.
Calculate the capacitances of A and B.
{1. 150 V, 3 mC,2. 45 μF, 4.615 μF,3. 3.6 μF, 72 μC, 288 μJ,4. 1200 μC; 120 V, 80 V; 6 μF,5. 500
μC; 250 V, 167 V, 83 V; 0.0208 J,6. 15 μF in series,7. 2.77 μF, 18.46 μC,8. 4.57 μF, 3.56 μF,9.
3.6 μF, 5.4 μF}

102
5.7 Permittivity
At any point in an electric field, the electric field strength E maintains the electric flux and produces
a particular value of electric flux density D at that point. For a field established in vacuum (or for

practical purposes in air), the ratio D/E is a constant ε 0 , i.e.

where ε0 is called the permittivity of free space or the free space constant.

The value of ε0 is 8.85 × 10−12 F/m.


When an insulating medium, such as mica, paper, plastic or ceramic, is introduced into the region
of an electric field the ratio of D/E is modified:

where εr , the relative permittivity of the insulating material, indicates its insulating power
compared with that of vacuum:
5.8 Relative Permittivity

The product ε0εr is called the absolute permittivity,


ε, i.e.
ε = ε0εr
The insulating medium separating charged surfaces is called a dielectric. Compared with
conductors, dielectric materials have very high resistivities. They are there- fore used to separate
conductors at different potentials, such as capacitor plates or electric power lines.

103
5.9 The parallel plate capacitor

Figure 5.8

For a parallel-plate capacitor, as shown in Fig5.8.(a), experiments show that capacitance C is


proportional to the area A of a plate, inversely proportional to the plate spacing d (i.e. the dielectric
thickness) and depends on the nature of the dielectric:

Another method used to increase the capacitance is to interleave several plates en plates are shown,
forming nine capacitors with a capacitance nine times that of one pair of plates. If such an
arrangement has n plates then capacitance
C ∝ (n − 1). Thus capacitance

104
Example
(a) A ceramic capacitor has an effective plate area of 4 cm2 separated by 0.1 mm of ceramic of
relative permittivity 100. Calculate the capacitance of the capacitor in picofarads.
(b) If the capacitor in part (a) is given a charge of 1.2 μC what will be the p.d. between the
Plates?
Solution

5.10 Charging and discharging of capacitor


When the capacitor is fully discharged and the switch connected to the capacitor has just been
moved to position A as shown in figure 5.9. The voltage across the 100uf capacitor is zero at this
point and a charging current ( i ) begins to flow charging up the capacitor until the voltage across
the plates is equal to the 12v supply voltage. The charging current stops flowing and the capacitor
is said to be “fully-charged”. Then, Vc = Vs = 12v.

105
Figure 5.9

5.11 Uses of capacitor


A capacitor is a device used in a variety of electric circuits—for example, to tune the frequency
of radio receivers, eliminate sparking in automobile ignition systems, or store short-term energy
for rapid release in electronic flash units.
5.11.1 Camera Flash Attachments:
One practical device that uses a capacitor is the flash attachment on a camera. A battery is used to
charge the capacitor, and the stored charge is then released when the shutter-release button is
pressed to take a picture. The stored charge is delivered to a flash tube very quickly, illuminating
the subject at the instant more light is needed.
5.11.2 Computer Keyboards:
Computers make use of capacitors in many ways. For example, one type of computer keyboard
has capacitors at the bases of its keys, as in Figure 5.10. Each key is connected to a movable plate,
which represents one side of the capacitor; the fixed plate on the bottom of the keyboard represents
the other side of the capacitor. When a key is pressed, the capacitor spacing decreases, causing an
increase in capacitance. External electronic circuits recognize each key by the change in its
capacitance when it is pressed.

106
Figure 5.10 When the key of one type of keyboard is pressed, the capacitance of a parallel-plate
capacitor increases as the plate spacing decreases. The substance labeled “dielectric” is an
insulating material.
5.11.3 Electrostatic Confinement:
Capacitors are useful for storing a large amount of charge that needs to be delivered quickly. A
good example on the forefront of fusion research is electrostatic confinement. In this role,
capacitors discharge their electrons through a grid. The negatively charged electrons in the grid
draw positively charged particles to them and therefore to each other, causing some particles to
fuse and release energy in the process.
5.11.4 Defibrillators
In practice, there is a limit to the maximum energy (or charge) that can be stored in a capacitor. At
some point, the Coulomb forces between the charges on the plates become so strong that electrons
jump across the gap, discharging the capacitor. For this reason, capacitors are usually labeled with
a maximum operating voltage. (This physical fact can actually be exploited to yield a circuit with
a regularly blinking light). Large capacitors can store enough electrical energy to cause severe
burns or even death if they are discharged so that the flow of charge can pass through the heart.
Under the proper conditions, however, they can be used to sustain life by stopping cardiac
fibrillation in heart attack victims. When fibrillation occurs, the heart produces a rapid, irregular
pattern of beats. A fast discharge of electrical energy through the heart can return the organ to its
normal beat pattern. Emergency medical teams use portable defibrillators that contain batteries
capable of charging a capacitor to a high voltage. (The circuitry actually permits the capacitor to
be charged to a much higher voltage than the battery.) In this case and others (camera flash units
and lasers used for fusion experiments), capacitors serve as energy reservoirs that can be slowly
charged and then quickly discharged to provide large amounts of energy in a short pulse. The
107
stored electrical energy is released through the heart by conducting electrodes, called paddles,
placed on both sides of the victim’s chest. The paramedics must wait between applications of
electrical energy due to the time it takes for the capacitors to become fully charged. The high
voltage on the capacitor can be obtained from a low-voltage battery in a portable machine through
the phenomenon of electromagnetic induction
Practice question
1. How should three capacitors and two batteries be connected so that the capacitors will store the
maximum possible energy?
Answer
Explanation The energy stored in the capacitor is proportional to the capacitance and the square
of the potential difference, so we would like to maximize each of these quantities. If the three
capacitors are put in parallel, their capacitances add, and if the batteries are in series, their potential
differences, similarly, also add together.
2. Consider the combination of capacitors in Figure P. and P1
(a) What is the equivalent capacitance of the group?
(b) Determine the charge on each capacitor.

Figure p and P1

108

You might also like